You are on page 1of 344

WATER CRISIS: MYTH OR REALITY?

BALKEMA – Proceedings and Monographs


in Engineering, Water and Earth Sciences
Water Crisis: Myth or Reality?
Marcelino Botin Water Forum 2004

Edited by
Peter P. Rogers
Harvard University, Cambridge, Massachusetts, USA

M. Ramón Llamas
Royal Academy of Sciences, Madrid, Spain

Luis Martínez-Cortina
Spanish Geological Survey, IGME, Spain

London/Leiden/New York/Philadelphia/Singapore
Copyright © 2006 Taylor & Francis plc., London, UK

All rights reserved. No part of this publication or the information contained herein may be
reproduced, stored in a retrieval system, or transmitted in any form or by any means,
electronic, mechanical, by photocopying, recording or otherwise, without written prior
permission from the publishers.

Although all care is taken to ensure the integrity and quality of this publication and the
information herein, no responsibility is assumed by the publishers nor the authors for any
damage to property or persons as a result of operation or use of this publication
and/or the information contained herein.

Published by: Taylor & Francis/Balkema


P.O. Box 447, 2300 AK Leiden, The Netherlands
e-mail: Pub.NL@tandf.co.uk
www.balkema.nl, www.tandf.co.uk, www.crcpress.com

ISBN 10: 0-415-36438-8


ISBN 13: 9-78-0-415-36438-6

Printed in Great Britain


TABLE OF CONTENTS

PRESENTATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . VII
FOREWORD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . IX

I – WATER POLICY AND MANAGEMENT


1 Water governance, water security and water sustainability . . . . . . . . . . . . . . . . . . . . . . . . . . 3
P. Rogers
2 Water for growth and security . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
W.J. Cosgrove
3 Collective systems for water management: is the Tragedy of the
Commons a myth? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
E. Schlager & E. López-Gunn

II – THE ECONOMIC VALUE OF WATER


4 The economic conception of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
W.M. Hanemann
5 The value of water and theories of economic growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
M.S. Aguirre

III – IRRIGATION
6 Irrigation efficiency, a key issue: more crops per drop . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
K.D. Frederick
7 Irrigation efficiency, a key issue: more crops per drop.
Observations and comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
P. Arrojo

IV – VIRTUAL WATER
8 Virtual Water – Part of an invisible synergy that ameliorates water scarcity . . . . . . . . 131
J.A. Allan
9 Virtual Water – Part of an invisible synergy that ameliorates water scarcity:
commentary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
J. Ramirez-Vallejo
VI Table of Contents

V – GROUNDWATER
10 Significance of the Silent Revolution of intensive groundwater use in
world water policy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
M.R. Llamas & P. Martínez-Santos
11 Is intensive use of groundwater a solution to world’s water crisis? . . . . . . . . . . . . . . . . 181
A. Mukherji

VI – WATER AND POVERTY ALLEVIATION


12 Water and poverty alleviation: the role of investments and policy interventions . . . . . 197
R. Bhatia & M. Bhatia
13 Do investments and policy interventions reach the poorest of the poor? . . . . . . . . . . . 221
C.A. Sullivan

VII – WATER AND NATURE


14 Water and nature. The berth of life . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
F. García Novo & F. García Bouzas
15 Water and nature: a critical link for solving the water management crisis . . . . . . . . . . 253
G. Bergkamp

VIII – NEW TECHNOLOGIES TO COPE WITH WATER SCARCITY


16 Water recycling – A relevant solution? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
T. Asano
17 Urban and industrial watersheds and ecological sanitation: two sustainable
strategies for on-site urban water management . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
D. Del Porto
18 The potential for desalination technologies in meeting the water crisis . . . . . . . . . . . . 297
J. Uche, A. Valero & L. Serra
19 The potential for desalination technologies in meeting the water crisis: comments . . 323
E. Custodio

AUTHOR INDEX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333


PRESENTATION

Media headlines around the world often voice an impending water crisis that is likely to affect a
large share of the world population. A good number of experts consider that this looming crisis
might be over-hyped, and probably the consequence of a preventive strategy. However, crying
wolf seems to be backfiring: pessimistic predictions done in a good number of international
fora have not occurred and the credibility of a part of the scientific community is in jeopardy.
In an attempt to contribute to dimension the problem in a realistic way the Marcelino Botin
Foundation organized and sponsored the first Santander Workshop, entitled Water Crisis:
Myth or Reality? Topic and expert selection were entrusted to Prof. M. Ramon Llamas,
from the Complutense University of Madrid, Spain; and to Prof. Peter Rogers, from Harvard
University, USA.
A three-day workshop was held in the Marcelino Botin Foundation headquarters in
Santander from June 14 to 16, 2004. Two experts were assigned to discuss each of the topics.
The first one to present his/her views and the second to critique those opinions or to present a
different perspective on the same topic. A debate by the whole group followed. Experts were
required to circulate their manuscripts in advance and those wishing to do so were given time
to modify them for publication after the workshop.
This book includes a foreword and nineteen chapters corresponding to the topics selected.
Four of the manuscripts previously envisioned could not be finally included in the book for
different reasons. The task of collecting the final versions of the manuscripts from the authors
and to send them to the publishers has been painstakingly carried out by Dr. Martinez-Cortina.
The reader will find different opinions on the same topics, depending on the views of
each author. This is a consequence of the multifaceted character of most water-related issues.
However, all experts seem to have reached a general agreement on something: a potential

Figure 1. Headquarters of the Marcelino Botin Foundation in Santander (Spain).


VIII Presentation

Figure 2. Participants in the Santander Workshop1 .

water crisis would not be due to physical water scarcity but rather to water resources misman-
agement, or in other words, to poor water resources governance.
The Marcelino Botin Foundation has had prior involvement in dealing with water resources
issues, beginning in 1999, with a four and a half year research project on groundwater
resources: the Proyecto Aguas Subterraneas (PAS) also led by Prof. Llamas. The results
of this activity are summarized in thirteen monographs directly published by the M. Botin
Foundation (they can be downloaded from the Foundation Web: www.fundacionmbotin.org),
six books in Spanish published jointly with Mundi-Prensa, and two books in English published
by Balkema Publishers, as a common effort with the Geological Survey of Spain (IGME). A
basic Hydrogeology textbook for secondary school students has also been produced jointly
with IGME, and distributed by Spain’s Ministry of Education to all the country’s Secondary
Schools. This book has been translated into English for its inclusion among UNESCO’s publi-
cations. A final result of the PAS project was the organization of an International Symposium
on Groundwater Intensive Use in Valencia (December 10–14, 2002). Selected papers of this
Symposium gave rise to the second book published with Balkema.
Finally, we are glad to announce that the Marcelino Botin Foundation is already working on
the second Santander Workshop, to be held in June 2007, under the title of Ethical Issues in
Water Resources Management. Prof. Llamas will also be in charge of organizing this event.
UNESCO has already expressed a willingness to be involved.

Santander, September 2005


Marcelino Botin Foundation

1
From left to right; (standing): E. Custodio, T. Asano, P. Arrojo, D. Jiménez-Beltrán, J.A. Allan, D. Del
Porto, P. Rogers, M.R. Llamas, F. García-Novo, J. Ramírez-Vallejo, P. Martínez-Santos, W.M. Hanemann,
L. Martínez-Cortina, and W.J. Cosgrove; (seated): A. Mukherji, E. Schlager, C.A. Sullivan, M.S. Aguirre,
and E. López-Gunn.
FOREWORD

Everyday we are bombarded by the media with tales of gloom and doom – we are running
out of water, petroleum, open space, clean air, arable land, etc. While there is good reason
to be wary of some of the potential outcomes of these prognostications, the reader should
be encouraged to remember that forecasts are not destiny. The human race has an uncanny
knack of proving the doomsayers wrong. This certainly goes back at least as far as the 18th
century of Malthus and his gloomy predications of widespread famine unless there were wars
and epidemics to curb the natural growth of population. To be sure, there have been many
wars and epidemics since Malthus’s time, too many, nevertheless the population of the globe
continued its giddying increase. The outcome is that 200 years after Malthus, the population
has increased several-fold and enjoys greater longevity and health than it enjoyed in his times.
Of course, there are more poor people on the globe today than the total population at Malthus’s
time; that is bad, but there are also many more people who would be considered to be living
like kings in Malthus’s time. Is this good? Both poverty and affluence stress the environment
in ways that Malthus could never have envisaged. This book is about fresh water and asks the
question: is the impending global shortage of water reality or myth?
Water is a classic renewable resource, that is one that will cycle continuously regardless of
ice ages and global warming. It is pretty much in a fixed amount on our small planet. What is
at stake is the allocation of this amount of water among potential users. For example, before
homo sapiens roamed the globe, water use was 100% for the ecosystem. As mankind grew
and prospered we started to make inroads in this exclusive use of water by the ecosystem,
but here was plenty of water to go around for millennia. It took maybe a hundred thousand
years before homo sapiens demanded, and took, a significant amount of water from nature.
This was when agriculture became settled and we began to develop irrigated agriculture in
the Fertile Crescent and in China. Settled agriculture allowed a rapid increase in human
populations which continually demanded more and more water from nature’s account. Seven
thousand years brings us up to the present when the largest human demand for water comes
from irrigation systems worldwide. As much as 70–90% of all the water consumed on an
annual basis is used by irrigated agriculture. This is why so much of the running-out-of-water
stories are based upon the facts that in many parts of the world the rivers are running dry
and the groundwater is over-pumped to provide irrigation for feeding the world’s population.
Surely, this is a sign of a Global Water Crisis? Because if we are unable to feed ourselves, then
Malthusian constraints will take over with disastrous consequences for homo sapiens and for
the planet.
In addition, the rate at which water sources are becoming contaminated with waste from
humans, industry, and agriculture is truly alarming. Human ingenuity is busily creating tens
of thousands of new chemicals that ultimately find their way into our drinking water supplies.
More than one million three hundred thousand children under the age of four die each year due
to diarrhoeal diseases and another one million due to malaria, both largely caused by water
supply and management deficiencies. Global climate change is predicted by some to have
major impacts upon the availability, spatial distribution, and the variability of water supplies
during the next century.
This litany of disasters can make even the most stouthearted falter. This volume attempts to
examine these issues and tries to unravel from them what is permanent and unchangeable and
X Foreword

what is remediable and changeable. On close examination of the problems we find that there
is much to be hopeful about the global water situation. To be sure, there will be places on the
globe that will be condemned to permanent water shortages and crises, but these are only a
small part of the problems. We find that if the policy prescriptions outlined in these chapters
are taken into account, the Global Water Crisis can be dealt with over a fairly short time scale.

September 2005
The Editors
I

Water Policy and Management


CHAPTER 1

Water governance, water security and water sustainability

P. Rogers
Harvard University, Cambridge, Massachusetts, USA

ABSTRACT: This chapter attempts to bring together a set of disparate concepts that are fundamental
to examining water as a resource and establishing the seriousness of the current and future water scarcity.
As is well known, there is a plentiful supply of water considered at a global scale. However, as we examine
scales much closer to individual humans, a pattern of great heterogeneity emerges. Some parts of the
world have plentiful supplies of water, others have severe droughts; some plenty of high quality water,
others with badly polluted waters; in some the rivers flow full, in others they are devoid of water for many
days of the year. It would be simple if these differences were due only to the physical climate, but careful
examination shows that there are large differences within the same climate zones that cannot be explained
purely by climate and topography. In these cases, one sees the hands of human interference in terms of
governance, property rights, and sheer population size. The situation may become much more serious
in the badly impacted areas as the great climate change experiment unfolds. There is huge uncertainly
associated with the predictions of climate for the 21st century. The chapter is able to be optimistic in the
face of such uncertainty by pointing to several technical, economic, and social developments that can
reduce the human footprint on the scarce supplies of easily accessible water. By relying more on rainfed
agriculture and agricultural trade to meet food needs scarce irrigation water can focus upon higher value
and less water using crops or can be diverted to high value municipal and industrial uses; improving the
efficiency of current irrigation technologies will free up large quantities of water for other uses; relying
upon new ecological sanitation techniques can greatly reduce the impact on water quality; and low cost
breakthroughs of desalination cost which are now economically competitive with alternative sources of
fresh water to meet the needs of urban populations anywhere in the world. In order for these solutions to
the emerging crisis to be adopted much more attention will have to be paid as to how we as individuals
and communities approach and the world community approaches the governance of water. A successful
shift to effective governance will enable us to have sustainable water supplies for all well through the
21st century.

Keywords: Governance, policy, politics, water security, global resource availability, promising
options

“Of all the social and natural crises we humans face, the water crisis is the one that lies at
the heart of our survival and that of planet Earth . . . No region will be spared from the impact
of this crisis which touches every facet of life, from the health of children to the ability of
nations to secure food for their citizens . . . Water supplies are falling while the demand is
dramatically growing at an unsustainable rate. Over the next twenty years the average supply
of water worldwide per person is expected to drop by one third” (UN/WWAP, 2003).

1 INTRODUCTION

Writing about myths, reality, and crises, in the context of water governance, security, and
sustainability provides the writer with a huge number of studies on resource use by humans and
4 P. Rogers

nature. When considering sustainability, one has to contend with the 18th century Malthusian
hypotheses of fixed resources facing inexorable geometric human population growth. This
has to be reconciled with the 21st century’s actual water resource picture and the needs of
nature and ecosystems. Security of supply is a major issue, where one is faced with highly
variable resources having to meet the needs of increasing and highly diverse populations in
cities and countryside, both rich and poor, and the demands of nature. Finally, governance
has to face up to reconciliation of all these demands placed upon a fixed, or maybe even
a shrinking resource base. How can we ever unscramble the myths and realities from these
highly interrelated issues?

2 SUSTAINABILITY

The problem with the concept of sustainability is that everyone seems to understand what is
meant by it in its promiscuous usage, however, everybody understands it differently from all
others. As long ago as 1992, Pezzey (1992) compiled a list of over 50 definitions in widespread
use at that time. No two were identical, yet the users seemed to feel comfortable with their
use. Since then, a widespread outpouring of sustainable development literature triggered by
the UN Commission on Sustainable Development (UNCSD) has led to the addition of several
hundreds of indicators of sustainable development.
There are some deep philosophical issues involved in the concept of sustainability. First, we
need to understand sustainability for whom – the planet, the ecosystem and nature, the human
populations, the standard of living of the industrialized countries, the poor, the rich, for cities,
for suburbs, for particular species of fauna and flora, and deposits of metals and minerals, etc.
Second, sustainability is not a static concept; whatever we do now will influence, positively or
negatively, how sustainable the future course of human development will be. Hence, the time
horizon is of great importance. The Bruntland Commission (World Commission on Envir-
onment and Development, 1987) claimed that sustainability meant a process that “meets the
needs of the present without compromising the ability of future generations to meet their own
needs”. Repetto (1986) was more restrictive than the Bruntland Commission when he defined it
thus: “sustainable development as a goal rejects policies and practices that support current liv-
ing standards by depleting the resource base, including natural resources, and that leaves future
generations with poorer prospects and greater risks than our own”. But do we mean the next
generation or all future generations? The choice might change our conclusions about the out-
come. In addition, human populations and the ecosystem itself are quite adaptable to changes
in their resource base over time. We only have to examine the geological history of the recent
past when homo sapiens first showed up on the planet to see how adaptable our species and the
planet itself have been. Solow (1991) used a gentler definition: “leave future generations the
option or the capacity to be as well off as we are”. Pezzey (1992) makes the definition more
complex by raising the distinction between survivability “which requires welfare to be above a
particular threshold in all time periods” and sustainability “which requires welfare to be non-
decreasing in all time periods”. So without any agreed upon operational definition it is difficult
to say what in particular is sustainable – but we can unequivocally state what is not sustainable!
Our current flagrant over exploitation of energy and water resources is not sustainable.
Our concern in this book is water as a resource and hence we are spared some of the
anguish of persons working in broader ecosystems. Nevertheless, without water there are no
Water governance, water security and water sustainability 5

ecosystems. It is worth noting that sustainability also implies survivability, but that the concepts
are markedly different. Survivability implies that we will be able to survive particular episodes
of extreme stress whereas; sustainability has the implications of being able to sustain a life-
style at least at present levels well into the future. In the same vein, Pezzey (2002) further
confuses the debate as he points out how sustainability policy and environmental policy, usually
considered one and the same thing, can be theoretically quite distinct.

“Environmental policy is dynamic, government intervention to maximize inter-temporal social


welfare based on the individual’s own discount rate path, by internalizing the social values of
stocks and flows that agents ignore (externalize) when they privately maximize welfare . . . By
contrast, sustainability policy aims to achieve some improvement of intergenerational equity,
whether a general shift to a lower path of the utility discount rate over time, or a specific
goal such as making utility forever constant, or non-declining or sustainable” (Pezzey, 2002:
26–27).

In this chapter, then, the reader will notice the concern for managing to survive the next
two generations until 2050. In this stance we are more akin to the supporters of survivability
than of true sustainability. There are, of course, some ambiguities: should effects of climate
change be considered as sustainability issues or security issues?

2.1 Basic laws of ecology


There are three basic laws of ecology that must be obeyed if we are to achieve a semblance of
sustainability to our human existence on the planet Earth. These can be characterized in plain
language as:

(1) Everything must go somewhere;


(2) There is no such thing as a free lunch;
(3) Plus ça change plus la même chose.

The first law of ecology is recognized as the first law of thermodynamics; in other words
the conservation of matter and energy. First law violations are fairly obvious and direct.
They confront you immediately, such as when the solid waste that you throw out of the front
door soon clogs up the street in front of you. Either, you have to take remedial action, or
move your household fairly soon. It also reminds us that when one takes something from one
phase of the environment for use in another phase the first phase suffers a loss. For example,
diverting water from a river for domestic water supply removes water supply from aquatic
species.
The second law of ecology, which holds both in economics and in the natural sciences, is
recognized as the second law of thermodynamics. The second law is subtler. In the natural
sciences it applies where resources are used inefficiently; for example every energy trans-
formation increases the entropy of the globe. It speaks to species extinction in ecology, and
in economics it reminds us that every transaction has some external non-priced effects that
impose a cost upon third parties in addition to the costs borne by the direct participants in the
transaction.
The third law of ecology reflects human and social decision making in as much as things
change in the environment nations and states react pretty much the same way but with a time
lag. Diamond (2005) shows how culturally determined responses to environmental stress tie
6 P. Rogers

societies to outmoded practices which can lead to rapid social and environmental collapse. The
third law is based upon the nature of human societies and their perceptions of risks and the costs
of action, and tells us that developed societies and less developed societies behave in a similar
manner when faced with environmental and resource sustainability issues. So for instance, the
wealthy countries tolerated massive air and water pollution for decades before the perception
of the health and ecosystem damages became too intolerable and the major risks associated
with polluted drinking water were dealt with first, followed by clean-up of particulates in urban
air, then concerns for the ambient aquatic system were experienced and massive wastewater
clean-ups of the ambient water were started, shortly thereafter, the oxides of sulfur in the air
became of serious concern leading to the introduction of sulfur scrubbers on smokestacks,
and more recently, the issue of automobile emissions and other volatile compounds in the air
have become the targets of increasingly stringent controls. Finally, the transboundary issues of
long-range sulfur transport, ozone depletion, and carbon dioxide build-up in the stratosphere
have become a major concern of all countries, both rich and poor alike.
This sequence is exactly what we observe in the major developing countries: first a concern
for drinking water quality, then for particulates in urban air, followed by concerns for ambient
water quality protecting fish and marine species, the removal of sulfur from the air, the whole
urban mobile air quality improvement, and finally a concern for the greenhouse gas problems
caused by oxides of carbon and other gases. What is most interesting is the rapid rate that the
developing world is transitioning through this sequence in comparison to how long it took the
industrialized countries to make the same transitions. All too often we are impatient with
the rates at which change happens in the Third World, but we should bear in mind the distance
traveled in a span of decades rather than centuries that the original industrialized countries
took. The evidence (Shafik & Bandyopadhyay, 1992) seems to indicate an inverse U-curve,
the so-called Kuznets’ Curve, relating environmental insult to per capita income. In many
cases the argument implies that we can spend our way out of environmental degradation as
we become richer.
These three laws of ecology basically constrain what is achievable on planet Earth: we have
to keep track of all material flows; we have to avoid unpleasant irreversible effects; and we can
expect that most human societies, given time and sufficient wealth, will choose environmen-
tally sustainable paths – if it is not too late. The first two laws are immutable and not easily
subverted by social and economic tinkering – in the long run the second law of thermodynam-
ics guarantees us a heat death, although the speed with which we approach that final extinction
is certainly a social choice. The rate at which the third law guarantees a safer path towards
sustainability is entirely dependent upon social, political, and economic forces. In comparing
several societies’ radically different responses to environmental stresses (particularly increas-
ing aridity) Diamond (2005) clearly demonstrates that the outcomes can be radically different
depending upon the nature of the environmental threats and socio-cultural rigidities.
In this chapter we choose a pragmatic stance and use operational definitions when possible.
Therefore, we work on the principle (attributed to Solow, 1991) that if we can manage to be
sustainable for the next couple of generations, say until 2050, we will be able to provide our
descendents with the tools to be able to extend sustainability a further couple of generations,
and so on. It is important to remind ourselves that no species comes with a written guarantee
against extinction, least of all a dominating species like homo sapiens. A 10,000-year-in-the-
future look back to the present would be helpful, but we will never have such a luxury. We
have only our recent history and our brains to guess at such a future.
Water governance, water security and water sustainability 7

3 GLOBAL WATER: SHORTAGE, SCARCITY, AND STRESS

The reason for this book is to explore the notion that there is a global water crisis and to examine
whether it is more of a myth than a reality, or if a reality to suggest paths to reduce the magni-
tude of the crisis. To address these questions one needs to understand precisely where we stand
at present with respect to water availability and use, in order to make any speculation about
the future. As mentioned earlier, there is no universally acknowledged reliable global database
for water availability and use. The works of Russian geographers, L’Vovich (1979) and
Shiklomanov (2000) are considered to be fundamental, and are the usual starting point for the
discussions concerning global water use and availability. They have attempted the Herculean
task of estimating water availability and use in countries and regions around the globe. These
data have been the source of much speculation as to the capacity of the earth’s water sup-
plies to sustain future populations, for example, L’Vovich’s work lead directly to Falkenmark’s
(Falkenmark & Widerstrand, 1992) concern with regional water stress, now and in the future.
In order to discuss sustainability of the resource in the future, even if we have a good
estimate of current conditions, we still need to rely upon some sort of model to predict the future
conditions. At this point we run into the classical Malthusian problem; a growing population
meets a fixed resource. Obviously if we carry on using a particular resource the way we have
been without changing, then an increased population leads inexorably to a crisis in the amounts
of the resource available for consumption. When that will occur will depend upon many social,
technical, and economic choices. Recall that Malthus’ideas were first propounded in 1783 and,
despite a six-fold increase in the population since then, none of the predicted population-based
resource crises have occurred. The reasons for the failure of prognoses based upon such an
eminently reasonable proposal, and supported by unassailable logic, are many and lead us far
beyond the scope of this chapter. Suffice it to say that as scarcity begins to be felt, the economic
relationships of the resource use to other substitutable resources changes in such a way as to
reduce the demand for the embattled resource. For example, there is no reason to expect the
global population to keep on growing after it plateaus out at around 9000 millions in 2050. In
fact there is every reason to expect it to start to decrease, as currently in Europe, with increasing
standards of living. And if we remove the population growth Malthus no longer holds sway!
In addition, rapid technological developments have improved the efficiency of water use and
recycling so much that pushing Malthus into the future is much easier. However, there will
be a long lag time between the declining population and declining water use per capita. The
reason for this is that the levels of water use in many parts of the world (the majority of the
world’s population in fact) are currently so low that any improvement in standard of living
will necessitate large increases in per capita water use, despite improved technological fixes.
The recent United Nations/World Water Assessment Programme (UN/WWAP, 2003), Water
for People, Water for Life, gives the current best estimates of water and its uses and concludes
that we are. . .
“ … facing a serious water crisis. All of the signs suggest that it is getting worse and will
continue to do so, unless corrective action is taken. This crisis is one of water governance,
essentially caused by the ways in which we mismanage water” (UN/WWAP, 2003: 4).

It is important to note that the UN/WWAP study concludes that the crisis faced is a
governance crisis not a resource crisis. They are concerned about a water crisis, but one
that can be remedied by management and governance; not the classic Malthusian end point!
8 P. Rogers

As mentioned above, we have chosen to use the water availability data from Shiklomanov
(2000) in the sketches of global water stocks and flows. In Figure 1 we sketch the total water
cycle showing both stocks (shown in square boxes) and annual flows (shown by arrows). All of
the numbers are in thousand km3 . As expected there is typically a huge difference between the
stocks and annual flows with one major and surprising exception – the atmosphere. The stock
of water in the atmosphere at any one time is only about 12,900 km3 . Given that the annual
precipitation is on the order of 577,000 km3 , the retention time of moisture in the atmosphere
must be only in the order of 8 to 9 days! Glaciers and groundwater are the major terrestrial
reservoirs of fresh water on the globe. Of the 187,900 km3 in surface storages only about
8000 km3 are due to man-made storage reservoirs. Rivers also have very little storage in
comparison to annual runoff. In Figure 2 we rely upon a slightly different database used by
Postel et al. (1996), which estimates the fate of the renewable fresh water (RFWS in the sketch)
falling on land surfaces.
This diagram shows the division of the available fresh water into Falkenmark’s blue (surface
runoff; shown here in the medium grey shade on the right hand side of the figure), green (water
that falls to the surface and does not contribute to runoff; shown in light shading on the left
side), and brown (water diverted for human use and which is contaminated before being

Figure 1. Schematic global water balance.


Water governance, water security and water sustainability 9

returned; darker shade on the lower right). The important points of this figure are that more
water evaporates from the land surface than flows over or under it. Of this green water humans
are able to capture 18,200 km3 for rainfed crops, forest products, etc. Following the blue water
side of the diagram we see that the amounts of water appropriated for human use is less than
that appropriated from the green water. Of the total brown water (6780 km3 ) indicated in the
figure, at least 3000 km3 should be removed, as this is the estimated amount of additional
evaporation due to surface water irrigation. Hence, the total amount of brown water to be
dealt with as human pollutants is on the order of 3700 km3 . According to this figure, the
human appropriation of accessible renewable fresh water is on the order of 30% of the total.

RFWSland
(110,300 km3/yr)

Total runoff
Total
(40,700 km3/yr)
evapotranspiration
on land
(69,600 km3/yr) Remote flow
(7774 km3/yr)

Uncaptured
floodwater
(20,426 km3/yr)

Geographically and
Temporally accessible
runoff (AR)
(12,500 km3/yr)

Withdrawals Instream uses


[4430 km3/year (35%)] [2350 km3/year (19%)]

Human appropriation Human appropriation


of ET of AR
[18,200 km3/yr (26%)] [6780 km3/yr (54%)]

Human appropriation
of accessible RFWSland
[24,980 km3/yr (30%)]

Human appropriation
of total RFWSland
[24,980 km3/yr (23%)]

Figure 2. Annual terrestrial renewable fresh water supplies (Postel et al., 1996).
10 P. Rogers

Figure 3. Environmental water scarcity index by basin (World Resources Institute, 2003).

The questions that arise are: whether this is a large amount now, how large will it get in the
future, and will it be sustainable?
If we assume that we would need a 17% increase in the water diverted to irrigated agriculture,
based upon IWMI (2000), then the numbers in Figure 2 indicate that globally on the average
there is still a large margin of water available. There are, however, large regional variations in
water availability for agriculture. But if we look towards rainfed agriculture there is a huge
amount of green water evaporation that could be additionally appropriated for agriculture
in several productive agricultural regions. This may imply, however, the necessity of large
transshipments of grains and other foodstuffs from the rainfed regions of the globe to the less
well-watered regions. The option of relying on virtual water will always be available, as a
backstopping technology should more local solutions fail. To achieve this, a transformation in
many arid country’s policy on food self-sufficiency to food security would be needed. Figure 3
shows the estimates of current levels of scarcity by major river basins worldwide (World
Resources Institute, 2003).
The recycling of irrigation water evapotranspiration is an interesting question. Water vapor
has only a residence time of 8 to 9 days in the atmosphere. This may not seem to be a
long time, but given the circulation in the troposphere the additional water vapor typically
has moved hundreds of kilometers from its origin and reappears as precipitation in those new
locations mostly over the oceans. In terms of water balances carried out on an annual basis, the
evapotranspiration is already accounted for in the other locations. It should be remembered that
virtually no water is lost (or gained) annually on planet Earth. What we see is a reapportionment
of the location of the water (and water vapor) between fresh surface water, groundwater, and
the oceans with the atmosphere being one of the temporary channels to do this. When we talk
about consumptive use of water, we are, hence, referring to it as measured from the perspective
of an annual water balance. So irrigation return flows from crop drainage are available almost
immediately whereas the evapotranspiration component goes out of the immediate system and
is already accounted in the precipitation in other regions.
Water governance, water security and water sustainability 11

3.1 Security
One of the basic needs for humans is security. Before anything else, humans need security
of food and shelter, and security of personal freedom. Security of possessions, and life-style
come a distant third in this ranking. When we talk about water, however, it is not just the
cup of water today that humans’ need, it is the ensured supply and access well into the future,
measured in terms of the life expectation of the individuals concerned. Security of supply
then, concerns access and property rights that allow different groups to satisfy their needs.
One additional wrinkle with water security is that not only is a shortage of water (drought)
a security risk, but also a surplus of water (flood) is a security risk. In many parts of the
world floods follow droughts leading to large-scale disruptions of human activities and loss
of life. So, humans need security against droughts and floods. Moreover, with the widespread
pollution of the surface and ground waters, humans need security from contamination of the
supply. Since the terrorist attacks of September 11, 2001, security in the USA has focused
more on dealing with willful terrorist acts rather than broader acts of man and nature.
There are many concepts floating around water security that are often used synonymously
with water security. Water stress and water scarcity, water vulnerability, water shortage, and
basic human needs appear in the literature when various aspects of water security are under
discussion. Winpenny (pers. comm.) expanded on some of these concepts and gave examples
of their use in managing water security:
“In popular usage, scarcity is a situation where there is insufficient water to satisfy normal
requirements. However, this commonsense definition is of little use to policy makers and plan-
ners. There are degrees of scarcity – absolute, life-threatening, seasonal, temporary, cyclical,
etc. Populations with normally high levels of consumption may experience temporary scarcity
more keenly than other societies, who are accustomed to using much less water. Scarcity often
arises because of socio-economic trends having little to do with basic needs. Defining scarcity
for policy-making purposes is very difficult.
(1) Water shortage: a dearth, or absolute shortage; low levels of water supply relative to
minimum levels necessary for basic needs. Can be measured by annual renewable flows
(in m3 ) per head of population, or its reciprocal, viz. the number of people dependent on
each unit of water (e.g. millions of people per km3 ).
(2) Water scarcity: an imbalance of supply and demand under prevailing institutional arrange-
ments and/or prices; an excess of demand over available supply; a high rate of utilization
compared to available supply, especially if the remaining supply potentials is too difficult
or costly to tap. Because this is a relative concept, it is difficult to capture in single indices.
However, current utilization as a percentage of total available resources can illustrate the
scale of the problem and the latitude for policy makers.
(3) Water stress: the symptoms of water scarcity or shortage, e.g. growing conflict between
users and competition for water, declining standards of reliability and service, harvest
failures and food insecurity. Difficult to capture in numbers, though a checklist approach
is possible” (Winpenny, pers. comm.).

For absolute levels of these definitions Population Action International (2003) gave the
following:
“A country whose renewable fresh water availability, on an annual per capita basis, exceeds
about 1700 m3 will suffer only occasional or local water problems. Below this threshold coun-
tries begin to experience periodic or regular water stress. When fresh water availability falls
below 1000 m3 /yr per person, countries experience chronic water scarcity, in which the lack
12 P. Rogers

of water begins to hamper economic development and human health and well-being. When
renewable fresh water supplies fall below 500 m3 per person, countries experience absolute
scarcity” (Population Action International, 2003).

A problem with these definitions is that they do not distinguish between security concerns
due to the actions of humankind or of nature. The issue of potential climate change is a real
test for the need to make sharp distinctions. Many climate change effects appear to be purely
natural events unfolding on a geological time frame; others appear to be human induced. How
do we separate these effects? Should we? If it is an issue of control we have little choice; we
can only influence the human-induced effects and work out strategies to mitigate the natural
effects. This dichotomy is referred by the Intergovernmental Panel on Climate Change, Third
Assessment Report (IPCC, 2001) as:
(1) “Climate change will lead to an intensification of the global hydrological cycle and can
have major impacts on regional water resources, affecting both ground and surface water supply
for domestic and industrial uses, irrigation, hydropower generation, navigation, in-stream
ecosystems and water-based recreation”.
(2) “The impacts of climate change will depend on the baseline condition of the water supply
system and the ability of water resource managers to respond not only to climate change but
also to population growth and changes in demands, technology, and economic, social and
legislative conditions” (IPCC, 2001).
The first quote deals with climate change from all sources, the second addresses the ability
to control human causes and adjust to the natural causes. This is essentially what water profes-
sionals do in the normal course of their professional careers. From the climate data on the recent
past it is very difficult to discern trends in water availability and the model results reported in
the IPCC (2001) are even more difficult to interpret. For example in Figure 4 based upon the
IPCC, which compares two of the IPCC’s scenarios: A2 which refers to a scenario with a slow
decrease in fertility rates and regional patterns, and B2 that has more rapid fertility declines and
with local rather than regional economic growth rates. These are both middle of the road scenar-
ios, neither is at the extremes of the projection ranges. Nonetheless, Figure 4 shows wide and
confusing differences between the forecast made by the IPCC models for the middle of the cen-
tury. Agreement as to sign and magnitude of changes in precipitation mainly occur at the poles,
with mostly inconsistent results in the mid-latitudes. It is hard to draw strong conclusions from
these simulations. Alcamo et al. (2000) using the IPCC models and data on river basins under
severe stress in 1995, predicted in Figure 5 critical regions in the world for 2025. The results
are not much different from what we already know about regions currently under water stress.
Given the conflicting views on the effects of climate change on water availability and use,
and given our knowledge of the rapidity of demographic, technological, economic, and social
impacts on water use, leads us to endorse the IPCC positions quoted above and focus upon
management of the resource in all of its dimensions. In other words, we believe that effective
governance of water is the one fundamental approach that we know that will enable us to
achieve our survivability goal by 2050.

3.2 Solving water security issues


3.2.1 Use more green water
Based upon Falkenmark’s conceptual breakthrough, now in book form (Falkenmark & Rock-
strom, 2004), of partitioning water into blue and green water, we now have a radical new look
Water governance, water security and water sustainability 13

Figure 4. Comparison of scenarios A2 and B2 of TAR (IPCC, 2001).

Figure 5. Critical regions in the world for 2025 (Alcamo et al., 2000).
14 P. Rogers

at the hydrological cycle. Traditionally water balances kept track only of the blue water. By
this is meant all of the water that appears as runoff to the streams and as groundwater recharge,
but neither the green water involved in supporting the non-irrigated lands of the globe, nor
the used water (brown water), appear directly in the balances. Using this new concept, we
have redrawn the Postel et al. (1996) water balances in Figure 2 to emphasize a global water
balance that truly reflects the roles of blue, green, and brown water for supporting humanity
and the ecosystem. Of the 110,300 km3 /yr Reusable Fresh Water Supply (RFWS in Figure 2)
available from precipitation on the land surface of the globe, only 40,700 km3 /yr goes to the
blue water account, and 69,600 to the green water account. Figure 2 shows the brown water;
at most this amounts to about 25% of the human withdrawals, or 1100 km3 /yr. The Figure also
shows that human appropriations of the green water were almost three times as large as those
of the blue water. The possibility of apportioning more of the green water for crop production
and reducing the demands on the blue water are active questions that could lead to dealing
more easily with feeding the future populations.
Paying more attention to rainfed agriculture research and to trade of agricultural commod-
ities between the rainfed areas and the more arid areas are two policies that immediately
suggest themselves in this context as potent solutions to feeding the world.

3.2.2 Exploit the asymmetries in water use


There are two ways of characterizing water use: withdrawal use, and consumption use. With-
drawals reflect the amounts of water removed from the water source for use and are the
amounts of water actually consumed or evaporated in the production process. For most uses
small amounts of water are actually consumed with the bulk returning, albeit in many cases
polluted, to the water source. Irrigation uses, however, typically evaporate more than 70% of
the withdrawn water. Industry, particularly heavy industry and fossil fuel electric generation,
withdraw huge amounts of water, but consume little. Municipal and commercial withdrawals
are typically the smallest withdrawals with about 20% consumptive use. This leads to the role
of irrigation water as a balancing mechanism for meeting water demands in other sectors.

3.2.3 The irrigation flywheel


In many countries of the world as much as 60–95% of all water is consumed by irrigated
agriculture. This fact is extremely important when looking toward future potential water short-
ages. Typically, a 10% reduction in irrigation water consumption would make available a more
than doubling of the amounts of water withdrawn for industrial and municipal uses. In other
words, provided that effective wastewater management programs are in place, the amounts
of water made available for municipalities and industries could be effectively doubled. The
question then hinges on whether or not such a 10% reduction in irrigation efficiencies could
be achieved (or equivalent areas for rainfed crops developed) at the same time as agriculture
is being expanded to meet increasing demands for food, fiber, and other industrial crops.

3.2.4 Virtual water escape hatch


Closely related to the irrigation flywheel is what Allan (2001) characterized as virtual water:
the use of imports of food crops as a substitute for use of domestic water for irrigation.
Essentially the importing country is importing the water that was used to grow the crops in
the exporting country. This virtual water can amount to as much as 1–5000 tons of water per
ton of crop imported. Hence, an import of 2 million tons of food from a rainfed source will
Water governance, water security and water sustainability 15

save the importer 2000–10,000 million tons (or 2000–10,000 Mm3 ; the annual average flow
of the Nile is approximately 60,000 Mm3 ) of domestic water. Virtual water, through the global
agricultural trade is already helping overcome the wide disparities between the distributions
of water resources among countries, and can be expanded to help future needs. Based upon
estimates of the world agricultural trade, the total virtual water already amounts to as much
as 700,000–800,000 Mm3 /yr (Hoekstra & Hung, 2002; Ramirez-Vallejo & Rogers, 2004).

3.2.5 Urban water: low cost desalination breakthrough


Recent developments in membrane filtration technologies combined with reverse osmosis
have revolutionized the potential for widespread application of desalination. Costs as low as
US$ 0.50/m3 (about US$ 1.80 per 1000 gallons) are expected for large water treatment plants.
At this cost, all urban areas in the world with access to saline water could have a plentiful water
supply comparable to current typical prices for urban water supply. While this does not give
an instant solution to the problem in many poor cities in the developing world, it does indicate
that reasonable economically achievable technology can solve the problems as they develop.

3.2.6 The ecological revolution


The major reason for widespread dissatisfaction about water issues in developing countries
stems from the vast numbers who have no access to adequate drinking water or sanitation.
The numbers often used are 1000 million without adequate water supply and 2000 million
without access to adequate sanitation. Whatever the actual numbers are, they represent a
massive failure by national governments and the world community to deal with these two
issues, which involve human dignity much more than economics. It is economics, however,
which drives the lack of action in meeting the water needs. It is simply beyond the financial
capabilities of existing governmental and international agencies to cover the capital cost of
providing conventional water and sanitation to these persons and their increasing numbers.
Either large new funding sources will be needed, or the conventional views of water-borne
sanitation will have to be changed toward well-known ecological sanitation practices that use
little or no water. Low cost ecological engineering breakthroughs are likely to revolutionize
the supply of water and sanitation in the near future.

3.2.7 Resolving transboundary conflicts


Two or more riparian countries share more than 261 river basins worldwide, and 40% of the
global population lives within these basins (Wolf et al., 2003). Often the basins are shared by
water-rich and water-poor countries or by heavily polluting industries in some countries and not
in others. Since there is no strong international law governing the resolution of transboundary
water quality and quantity disputes, the potential for physical and armed conflict remains high.
This issue needs careful consideration by riparians and the regional family of nations to head
off these conflicts from becoming worse. If these issues could be resolved, for example, 500
million inhabitants in the Ganges-Brahmaputra basin could improve their sense of security
about future water supplies and their access to them. Similar cases are exemplified in the Nile,
Mekong, and the Tigris-Euphrates basins.

3.2.8 Reform of idiosyncratic water institutions


Unfortunately, one water fact-of-life makes it very difficult to implement the types of tech-
nical and social breakthroughs already conceptually available: the hurdle is that property rights,
16 P. Rogers

institutions, and laws, vary widely from country to country even within countries. The plethora
of institutions and legal regimes in all countries tends to get in the way of rational water manage-
ment. In many cases it is impossible to reallocate or trade between conflicting water uses. The
confusion allows for the entrenchment of conventional institutional views and blocks consid-
eration of newer approaches. Approaches such as public/private partnerships are often difficult
to implement under these Balkanized conditions. Any progress in this dimension will make
the potential for implementing the above measures that will greatly enhance water security
for all.

4 WATER GOVERNANCE

At the start of the new millennium the concept governance has become part of the lexicon
of all development planners. The World Bank, the Asian Development Bank, and most UN
Agencies have embraced the concept. It also features large in the discussions of the bilateral
agencies and the NGO community. On closer examination, however, we can see that in most
cases it is merely a renaming of concern for institutions and institution building that goes a
long way back in time in most agencies. The concern for institutions, and now governance, has
grown out of the slow realization that the problems of development are not simply shortages of
capital. Indeed there were so many cases where development happened without large capital
transfers that some other, X-factor, was needed to explain them. The fads and fashions swung
from education to health, from poverty alleviation to globalization, etc. We now stand at a
period of time at which the role of governance – the relationship between the state and civil
society has been identified as the critical X-factor.
“Governance is the exercise of economic, political and administrative authority to manage
a country’s affairs at all levels . . . It comprises the mechanisms, processes and institutions
through which citizens and groups articulate their interests, exercise their legal rights, meet
their obligations and mediate their differences . . . Water governance refers to the range of
political, social, economic and administrative systems that are in place to develop and manage
water resources, and the delivery of water services, at different levels of society” (GWP, 2000).

Governance relates to the broad social system of governing, which includes, but is not
restricted to, the narrower perspective of government as the main decision-making political
entity. There is no single definition of governance that effectively includes all of these dis-
parate ideas and different approaches are followed. Some may see governance as essentially
preoccupied with questions of financial accountability and administrative efficiency. Others
may focus on broader political concerns related to democracy, human rights and participa-
tory processes. There are also those who look at governance with a focus on the match and
mismatch between the politico-administrative system and the ecological system or in terms of
operation and management of services. Water governance is already practiced in all countries
and our aim here is to make it more transparent and relate it directly to sustainable water
development and security.
Governance covers the manner in which allocative and regulatory politics are exercised in
the management of resources (natural, economic, and social) and broadly embraces the formal
and informal institutions. Institutions are interpreted here to include both the formal (codified
and legally adopted) and the informal (traditionally, locally agreed and non-codified). The
Water governance, water security and water sustainability 17

new term for discussing this combination of formal and informal institutions is distributed
governance. There is a profoundly political element to governance, which involves balancing
various interests and political realities that must be taken into account. Although politics may
set the agenda, the priorities and the vision, people need governance systems that give the
political vision credibility and ownership. Finally, management structures must be established
to run things and carry out the day-to-day tasks.
The notion of governance for water includes the ability to design public policies and insti-
tutional frameworks that are socially accepted and mobilize social resources in support of
them. Water policy and the process for its formulation must have as its goal the sustainable
development of water resources, and to make its implementation effective, the key actors/
stakeholders must be involved in the process. Governance aspects overlap with technical and
economic aspects of water, but governance points us to the political and administrative ele-
ments of solving a problem or exploiting an opportunity. Governance of water is a subset of
the more general issue of the creation of a nation’s physical and institutional infrastructure
and of the still more general issue of social cooperation.
Water governance is concerned with those political, social and economic organisations
and institutions (and their relationships), which are important for water development and
management. Given the complexities of water use within society, developing, allocating and
managing it equitably and efficiently and ensuring environmental sustainability requires that
the disparate voices are heard and respected in decisions over common waters and use of scarce
financial and human resources. Water governance is concerned with the functions, balances
and structures internal to the water sector (internal governance). It includes the framing
of social agreements on property rights and the structure to administer and enforce them
known as the law. Influences also come from civil society and from the current government
and these latter are all considered parts of the external governance of water, which will
be discussed later. Although issues can arise for water governance from the economic and
technical spheres, in most countries the driving force is politics. Effective governance of water
resources and water service delivery will require the combined commitment of government
and various groups in civil society, particularly at local/community levels, as well as the private
sector.
To achieve more effective water governance it is necessary to create an enabling environ-
ment, which facilitates efficient private and public sector initiatives and stakeholder involve-
ment in articulating needs.

4.1 Good governance matters


Social analysts have shown that there is a strong causal relationship between better governance
and better development outcomes such as higher per capita incomes, lower infant mortality
and higher literacy (Kaufmann et al., 1999). A stable and just social order founded on clear
institutional rules and effective and equitable markets enhances poverty reduction. Effective
governance is thus essential to poverty reduction and can help the poor to help themselves.
Poor governance is a barrier to development and hurts the poor through both economic and
non-economic channels, making them more vulnerable and unable to adapt to changes. As a
result, markets will be weak and distorted thus holding back growth and employment oppor-
tunities. Structural and institutional reforms are needed to turn poor governance into more
effective governance; including measures such as creating accountability in the use of public
18 P. Rogers

funds, building national capacity for better policy formulation, implementation, and enforce-
ment mechanisms. It includes making decision-making and implementation more inclusive
processes where civil society and the private sector have clear roles to play with shared responsi-
bilities on the basis of public-private partnerships. The division of labour between the different
actors and the sharing of responsibilities and balancing power relations are all part of the same
process, that of defining the governing system.

4.2 Historical context of water governance


Since the Dublin conference in 1992, significant international goals have been set that relate
to water governance. At the 2000 World Water Forum in The Hague, the GWP Framework for
Action (GWP, 2000) stated that “the water crisis is often a crisis of governance”, and identified
making water governance effective as one of the highest priorities for action. The 2000 Hague
Ministerial Declaration reinforced this view and called for “governing water wisely to ensure
good governance, so that the involvement of the public and the interests of all stakeholders
are included in the management of water resources”. At the Bonn 2001 Freshwater Confer-
ence the ministers recommended action in three areas, with water governance as the most
important. They proposed that “each country should have in place applicable arrangements
for the governance of water affairs at all levels and, where appropriate, accelerate water sector
reforms”. The UN 2000 Millennium Assembly, emphasised conservation and stewardship in
protecting our common environment and especially “to stop the unsustainable exploitation
of water resources, by developing water management strategies at the regional, national and
local levels, which promote both equitable access and adequate supplies”. This was endorsed
at the World Summit on Sustainable Development in 2002 where Heads of State agreed a
specific target “to prepare IWRM (Integrated Water Resources Management) and water effi-
ciency plans by 2005”. To be meaningful these plans will need to take cognisance of prevailing
governance systems and allow for necessary reforms.
At the beginning of the 21st century, we are searching for coherence and accountability
in the maze of organisations within national (and international) political systems. However,
many of today’s institutions and government systems were developed in the 19th century
to supervise states with much more limited functions than today. The developing coun-
tries face particular problems as they often have layers of systems – some indigenous and
others imported. It is not expected that developing countries can or even should adopt the
same systems as industrialised countries but there are basic principles for effective govern-
ance that they need to adopt in their own way. With regard to water, the state may need to
act quickly to develop the essential infrastructure for development and cannot wait for the
ideal governance systems to be established. Nevertheless, any water development should be
done hand in hand with broader governance reforms that will help to make the development
sustainable. In pushing for this care should be taken not to further weaken the weak state.
The traditional bases of political power have been eroded in the last 20 years or so and
the institutional strength of the state is being challenged. Some recent changes in society
have facilitated this weakening of the central state and making traditional water governance
irrelevant. Pierre (2000) claims that some of these changes (there are others) include:
• Fiscal crises within the state (limitations on raising taxes).
• The globalisation process, including deregulation of financial markets and volatility of
capital that restricts the state’s ability to govern/control the economy.
Water governance, water security and water sustainability 19

• Technological advances that facilitate networking and subsidiarity.


• A more assertive sub-national democracy in cities or semi-autonomous regions.
• Excessive workload and responsibilities on smaller government bureaucracies.
• Large concentrations of people and political power in urban areas.

There are several modes of governance that are often being pursued at the same time in
within a given state. In the following discussion we rely heavily on Pierre (2000) and Rogers &
Hall (2003).
Hierarchical governance: Part of modernisation is generally seen as the evolution of polit-
ical systems from top-down, hierarchical government systems with centralised institutional
settings, to more decentralised administrative forms. There is no evidence that more decen-
tralised systems are necessarily more effective than centralised ones. The real test here is, what
works in the particular setting. There is, however, a perceived ever-widening gap between those
countries that have managed to move toward subsidiarity – or the performance of functions at
the lowest effective level – and those that remain centralised and stagnant.
Market-led governance: With the end of the cold war in the closing decades of the 20th cen-
tury, many in the western countries proposed the market as the solution to economic growth,
social equity and environmental problems. This led to deregulation and more involvement
of the private sector and a changed role for the civil service and civil society. This institu-
tional restructuring of the state aimed to reduce government command and control functions
with more individualism (fewer collective solutions) and private enterprise and the market
as the superior resource allocation mechanisms. This market-led governance model is the
immediate background in which we now examine governance with respect to water resources
management and the delivery of water services. Today the honeymoon with the laissez-faire
market-led model is over, and hard questions are being asked. It is considered by many to be
too simplistic (hierarchies may not work well but markets do not necessarily work well either
in all situations) and not necessarily representative of wider societal values. More people are
examining what new instruments and new forms of exchange between state and society can
be developed to ensure political control and societal support. From this examination, propos-
itions for management in partnership, co-management and co-governance, and distributed
governance, have developed.
Distributed governance: At the beginning of the new millennium the state’s role of direct-
ing or steering society, particularly in the areas of environment and water, is being challenged
by cohesive local networks (civil society, private sector) and global networks (international
organisations and NGOs) with these same entities also supporting the state in its aims to
develop society. This gives a dynamic relationship between different social forces. Many
politicians (mainly in the West) see the state increasingly as part of the problem rather than the
solution. There are more calls for a return to smaller government, reversing the post Second
World War ideology of a hierarchical central state caring for its citizens. The state no longer
believes it can solve societal problems acting alone; particularly socio-environmental ones
and the private sector alone cannot address the problems of the poor and the environment.
The command and control or hierarchical model and the market-led governance models are
both thus much weakened. The Dublin Principles manifestly reflect this concept of distributed
governance.
Clearly modern governance sees formal authority being supplemented by an increasing
reliance on informal authority; for example, through genuine public-private co-ordination
20 P. Rogers

and co-operation to the benefit of both of these as well as the customer/citizen (organisations
such as the GWP (Global Water Partnership) and international NGOs such as Transparency
International are examples of such co-operative networks). The state thus needs to adapt
to a new situation and distributed governance is an institutional response to the changed
environment. Distributed governance is thus the empirical manifestations of state adaptation
to its external environment. It is the conceptual representation of the co-ordination of social
systems and specifically the role of the state in that process.

4.3 Water governance principles and legal bases


The Dublin Water Principles (WMO, 1992) bring water resources firmly under the state’s func-
tion of clarifying and maintaining a system of property rights, and through the principle of
participatory management at the lowest appropriate level asserts the relevance of meaningful
decentralization. The Dublin Principles that guide the IWRM approach are:
• Fresh water is a finite and vulnerable resource, essential to sustain life, development and
the environment.
• Water development and management should be based on a participatory approach, involving
users, planners and policy-makers at all levels.
• Women play a central role in the provision, management and safeguarding of water.
• Water has an economic value in all its competing uses and should be recognized as an
economic good.
There is increasing pressure to recognise and formalise water rights and this is happening
in many countries. Formalising rights raises complex questions about the plurality of claims
and the balancing of the distribution of benefits among the social groups. It also imposes
responsibilities including that of pollution prevention and financial sustainability. The process
of formalisation is often biased in favour of the rich and powerful who may abuse the system
and capture rights. Informal rights, as defined locally with their historical rules and principles,
are equally important and improper formalisation may lead to conflict between the formal and
traditional. The formalisation of rights may therefore be neither necessary nor sufficient to
secure access to water resources. The capacity to defend rights against competing claimants
is essential for the rights to be meaningful, whether they are formal or informal. An important
matter is to what extent the processes of devolving water rights serve segments of a population,
or its entirety.

4.4 Water law is about property rights


The state has an important role to play through its core function of defining property and use
rights and responsibilities. In modern pluralistic democratic societies, the foundation of the
state rests upon the publicisation (the term for the shift from the private to the public sphere,
also called nationalization in some countries) of the costly monitoring and policing needed
to protect productive assets from being redistributed to intruding claimants. Without this
policing, called the law, systems of property would never have advanced beyond appropriative
behaviour backed by force. Discussions of water rights usually focus upon the rights of the
property right holder and ignore the contingent responsibilities that holder has with regard
to others in society who do not share the rights. These obligations need to be stressed in any
discussion of governance.
Water governance, water security and water sustainability 21

The state has an important role to play in as much as it defines property rights. Some
examples of the main types of property rights and civil society’s responsibilities toward
different property rights with their associated rights and obligations:
• Open access property
There is no defined group of users or owners and the benefits are available to anyone. Indi-
viduals have both privilege (the ability to act without regard to the interests of others)
and no right (the incapacity to affect the actions of others) with respect to usage and
maintenance of the asset.
• Common pool property
The management group (the owners) has a right to exclude non-members and non-members
have a duty to abide by the exclusion. Individual members of the management group have
both rights and duties with respect to usage and maintenance of the property.
• Private property
Individuals have the right to undertake socially acceptable uses and a duty to refrain from
socially unacceptable uses. Others (non-owners) have a duty to allow socially acceptable
uses and a right to expect that only socially acceptable uses will occur.
• State property
Individuals have a duty to observe use and access rules determined by the controlling agency
of the state.
In most countries water is state property, with lesser amounts owned privately. However,
water property rights usually start out as an open access property, which is initially appropriated
by a group and becomes a common pool property resource. Ultimately the state tends to
appropriate these rights from the common-pool resource ownership group to create state
property. The state is then faced with the problem of how to deploy the resource to the national
advantage. Water rights and land use are closely linked, sometimes formally through riparian
rights, and landowners can affect water through land use changes such as deforestation. The
key to water governance at the beginning of the 21st century is how, through politics, states
can achieve this fairly and equitably, without reducing incentives for efficient use of the
resource.

4.5 Theories of collective action


The theoretical basis of governance lies in theories of collective behavior. Unfortunately, while
intellectually stimulating and of great historical interest going back to the Greek philosophers
and earlier, there is no one simple theory to explain every situation. Starting from an analysis
of property rights and experimentation with these rights over time has led the USA to a flexible
approach to water governance. This approach allows for shifts when the economic and social
conditions change without having to build institutions that cover all possible eventualities.
Many questions can be posed about the viability of any of the property rights regimes based
upon the collectivity of the individual players and their initial resource endowments. Robert
Wade (1987) cites Mancur Olson (1965) as the source of pessimism about the viability of
collective action in common pool or common property resources:
“One of the theories that have generated pessimism about the viability of collective action
is Mancur Olson’s logic of collective action (which might better be called the illogic of col-
lective action, or the logic of collective inaction). His core proposition is this: unless there
22 P. Rogers

is coercion or some other special device to make individuals act in their common interest,
rational, self-interested individuals will not act to achieve their common or group interests”
(Wade, 1987: 221).

Based upon his experiences with water management in Indian villages, Wade concludes
that this is not the case. Others, such as Ostrom (1990), again based upon Asian villagers’
collaboration on water allocation and use, back up Wade’s arguments. Wade and Ostrom
claim that groups can build mutually binding agreements in spite of substantial individual
differences, and that heterogeneity could facilitate cooperation when some members of the
privileged group value a collective good enough that they are willing to provide it in spite of
the actions or inaction of the remaining group members.
Tendler (1997) noted that we know a lot more about what constitutes bad governance than
we do about achieving good governance. Her case studies tend to question some conventional
nostrums and preconceptions of how governance should be and drive us back to a close func-
tional analysis of each individual case. Maass & Anderson (1978) provide in-depth analyses
of the development of the governance of irrigation since the 15th century in Valencia, Murcia,
and Alicante, in Spain. In all of these empirical studies the authors found strong evidence to
support the notion that, despite a wide range of property rights regimes, user groups could
develop into sustainable institutions over many years (centuries in the case of the Spanish irri-
gation property rights sharing systems). Essentially, there is a possibility of identifying a level
of centralisation and decentralisation and regulation to produce effective water governance.
Whilst empirical evidence suggests there can be no dogmatic solutions it would be helpful to
establish some universal attributes that make water governance effective in practice.

5 THE POLITICS OF WATER GOVERNANCE

The driving force in any area of governance is politics. The conventional view of the rela-
tionship of politics to governance portrays an orderly world where politicians act as rational
legislators in formulating laws for the general welfare, which in turn are implemented by
institutions, which carry out the legislated water policies through rules and regulations. Real
politics, however, is not so neat and in 1936 Harold Lasswell said it best merely in the title
of his book: Politics: Who gets What, When, and How? (Lasswell, 1936). Governance is now
not seen as a simple linear process, but as discursive and a highly complex set of interactions
between laws and institutions, and personal and group interests as well as the general interest.
Stone (2000: 208) says: “ . . . policy is more like an endless game of Monopoly than a bicycle
repair”.
For a water enterprise, politics is certainly part of the governance domain. Contestatory,
often very personalized maneuvering, aimed at the building of consensus and support for
policies and persons, is certainly part of the water politics concept. The application of poli-
tics to water problems has been called hydropolitics by John Waterbury in his cautionary tale
in the Hydropolitics of the Nile Valley (Waterbury, 1979), which demonstrated the powerful
combination of local and international politics. Hydropolitics can occur inside or outside a
water agency or enterprise, but the politics of water governance are primarily the play of the
sociological factors (structures, institutions, even leaders’ personalities) that lie outside the
water enterprise itself. These reflect more general political-sociology, that is to say, of
the water institution’s setting. Hence, framing the political decision is a fundamental question.
Water governance, water security and water sustainability 23

An agency’s own governance is nested within these factors and the boundary between the
agency and its environment is permeable, and social capital, which can take the form of
political power, can also move in both directions across it.

6 APPROACHES TO POLITICAL DECISION-MAKING

The literature on politics and political theory and their implications for effective water gov-
ernance leads to many different pathways for explaining political decision-making. A brief
introduction to them is given below.

6.1 Economic theory of politics


The literature on politics and political theory and their implications for effective water
governance in the USA leads to at least three different pathways for explaining political
decision-making. First, and a priori the most attractive to professional planners, is in the
direction of the Economic Theory of Politics (ETP), which is nicely summarized in a review
article by Jan-Peter Olters (2001). This theory aims at a synthesis of the ethics of Rawls with the
political theories of Lasswell, Dahl and Lindblom, and the welfare economics of Pareto, Hicks,
and Bergson. The ETP in essence attempts to replace the maximization of the Hicksian welfare
function with the maximization of a social welfare function. Downs (1957) led the attack on the
use of the classical economic social welfare function in ETP. Downs hypothesized that political
parties act in order to obtain income, prestige, and power, and politicians were motivated by the
vote maximization objective rather than by altruistic or ideological objectives, and concluded
that “parties formulate policies in order to be elected rather than win elections to formulate
policies”. Downs’ model is pure Adam Smith with selfish behavior of individual politicians
resulting in economic and other policies that guarantee a social optimum. In other words, “gov-
ernments continue spending until the marginal vote-gain from expenditure equals the marginal
vote-loss from financing” (Downs, 1957: 73). Despite Downs’s efforts, the ETP, however,
appears to be less relevant than other approaches to informing practical real political decisions.
An attempt to model water quality decisions using the ETP based upon Paretian Environmental
Analysis was made by Dorfman et al. (1972). Their rational approach provided the tools for
political bargaining, but did little to enlighten the bargaining process itself. Political solutions
do not necessarily rely upon Pareto feasibility to still claim to be good solutions.

6.2 Institutional approach


The second approach is an institutional approach exemplified by Maass et al. (1962) and
more recently by Crenson & Ginsberg (2002), which examines the evolution, development, and
erosion of popular democratic institutions and sees subtle shifts away from popular democracy
towards personal democracy. Popular democracy depends upon group political mobilizations
such as labor unions, fraternal associations, etc., whereas personal democracy has developed
asAmerican’s become increasingly atomized to small groups who exert their political influence
through political contributions leaving the political aspects to paid political operatives. This
pathway is very much in keeping with the Bowling Alone view of institutions of Putnam (2000).
The institutional approach assesses the institutions and sees how well they are meeting their
stated goals. Muddy Waters (Maass, 1951) used this approach in analyses of the relationships
24 P. Rogers

between the Congress and the Corps of Engineers. More recently, Crenson & Ginsberg (2002)
are particularly pessimistic about the general health of democratic institutions and democracy
in the USA. They see citizens becoming customers of the government, not the controllers
of the government. They give many reasons including the size of the population, the rise of
multitudinous NGOs, the lack of direct taxation specifically to fund government’s war-making
ability, and the role played by TV and media in converting citizens to passive consumers.
To give a sense of how much has changed in citizen participation in environmental and
water governance over the past decades consider the plethora of new citizen organizations that
have sprung up worldwide to help address these and other problems. Bornstein (2004) states:

“Consider that 20 years ago Indonesia had only one independent environmental organization.
Today it has more that 2000. In Bangladesh, most of the country’s development work is handled
by 20,000 NGOs: almost all of them were established in the past 25 years. India has well over
one million citizen organizations. Between 1988 and 1995, 100,000 citizens groups opened
shop in the former communist countries of Central Europe. In Canada, the number of registered
citizens groups has grown by more than 50% since 1987, reaching close to 200,000. In Brazil,
in the 1990s, the number of registered citizen organizations jumped from 250,000 to 400,000,
a 60% increase. In the USA, between 1989 and 1998, the number of public service groups
registered with the Internal Revenue Service jumped from 464,000 to 734,000, also a 60%
increase” (Bornstein, 2004).

It is difficult to get one’s head around the magnitudes involved. Even if they had only 50
active members in each group this would amount to more than 35 million active members of
these groups in the USA alone. Such participation levels would seem to contradict Crenson &
Ginsberg’s comments about the decline in participatory democracy in the USA.

6.3 The Polis model


The third, and most intriguing approach replaces the highly rational models of policy analysis
based upon ETP by the Polis model of Stone (2000). Stone argues that rational behavior
models (usually the ETP) miss the point of politics. She characterizes the rational behavior
model taught in schools of Public Policy as follows:
• Identify objectives.
• Identify alternative courses of action for achieving the objectives.
• Predict the possible consequences of each alternative.
• Evaluate the possible consequences of each alternative.
• Select the objective that maximizes the attainment of objects.
We can all recognize these steps as constituting most of what we do, or attempt to do, in our
professional and intellectual lives. They are the foundation of documents like the US Army’s
Corps of Engineers Principles and Guidelines (US Water Resources Council, 1983), which
would be unthinkable apart from the above criteria. So what could possibly be wrong with
such an approach? Stone’s response is:
“ . . . a model of political reasoning ought to account for the possibilities of changing one’s
objectives, of pursuing contradictory objects simultaneously, of winning by appearing to lose
and turning loss into victory, and most unusual, of attaining objectives by portraying oneself
of having attained them . . . Political reasoning is reasoning by metaphor and analogy” (Stone,
2000: 9).
Water governance, water security and water sustainability 25

In a series of tables she compares and contrasts the market model of policy analysis, which is
essentially based upon the ETP model, with her political (polis) model. For example in Table 1,
the comparison considers: what are the units of analysis, sources of conflicts, sources of ideas,
and what is the nature of collective action. From this table we see two radically different
perceptions emerging of the nature of society. The unit of analysis shifts from the individual
to the community. Self-interest is extended also to include public interest. The conflicts now
change from conflicts between self-interests to between self-interests and the public interest.
The nature of collective activity moves from competition to a mixture of cooperation and
competition. Most importantly, the criteria for decision-making change from maximizing self-
interest to the promotion of the public interest and loyalty to people and places. The nature
of the information used tends to move from objective to ambiguous, interpretive, incomplete,
and strategically manipulated. Ultimately, what matters under the Polis model are ideas, the
pursuit of power, and the maintenance of alliances. Polis approaches to politics and policy are
quite different from those implied by the rational model of politics.

6.4 The bureaucratic politics and process model


This model is based on political-bureaucratic bargaining in a federal system. Its focus is
typically the executive branch, with the elected legislature hardly in the picture. Classic cases
are drawn from USA foreign policy problems (for example, the Cuban missile crisis) where
Congress was not a major player, but this is the opposite from the situation in water, where
the executive branch until recently was largely excluded by Congress.

Table 1. Comparison of market model with polis model.

Concepts of society

Market model Polis model

1. Unit of analysis Individual Community


2. Motivations Self-interest Public interest (as well as self-interest)
3. Chief conflict Self-interest vs. self-interest Self-interest vs. public interest
(commons problems)
4. Source of people’s Self-generation within the Influences from outside
ideas and preferences individual
5. Nature of collective Competition Cooperation and competition
activity
6. Criteria for individual Maximizing self-interest, Loyalty (to people, places, organizations,
decision-making minimizing cost products), maximize self-interest,
promote public interest
7. Building blocks of Individuals Groups and organizations
social action
8. Nature of Accurate, complete, fully Ambiguous, interpretive, incomplete,
information available strategically manipulated
9. How things work Laws of matter (e.g. Laws of passion (e.g. human resources
material resources are are renewable and expand with use)
finite and diminish with use)
10. Sources of change Material exchange. Ideas, persuasion, alliances.
Quest to maximize own Pursuit of power, pursuit of own
welfare. welfare, pursuit of public interest.
26 P. Rogers

6.5 The congressional behavior model


A second federal model concentrates on the elected congress, with the view that to under-
stand congressional behavior is to understand that congressmen are single-minded seekers
of reelection. It follows from this that congressmen’s goals are to improve the welfare of
their constituents in the shortest possible time frame. The realities of information processing
are also important in describing congressional behavior. With humanly limited capacities to
absorb and judge, legislators are so overloaded with information that they have to be extremely
selective in committing their attention. Legislators deal with this by specializing in a particular
and limited area; in other domains they take their cues from other sources (colleagues, outside
groups, committee reports) that they have learned to trust.

6.6 The interest group model


When a national legislator thinks about the constituency that elected him, he or she rarely,
if ever, sees an undifferentiated mass of individual voters. They see categories of interests.
In some cases, they see only a few dominant interests. In the USA there are literally thou-
sands of active interest groups – environmental groups, water resources groups, professional
associations, and industry associations – involved with water policy (a total of 734,000 public
interest groups are registered with the Internal Revenue Service for all purposes including
water and environment). These interest groups often have overlapping concerns and overlap-
ping memberships. They constitute vital channels for particular publics to participate in the
federal governmental process. Pork barrel projects are the fodder for the well known iron
triangles of legislators, bureaucrats, and active interest groups that develop in specific issue
fields (the term pork barrel was first used to describe the acceptances of non-economic water
projects by congressmen in other districts in return for equally dubious projects in their own
districts in the bi-annual US Rivers and Harbors Acts).
Useful developments of interest group theory are found in Dahl’s (1961) regime theory and
in Buchanan & Tullock’s (1962) public choice theory. They attempt to predict which pattern of
decision-making will prevail based upon the concentration or diffusion of costs and benefits
of public choices. If we examine the distribution of effects of a particular action by the state,
depending upon whether the benefits (and the costs) are concentrated on a few persons or
widely dispersed throughout the economy, we can predict what type of political system will
dominate. Table 2 shows schematically how this can work. For example, it is often said that
the problem with water governance in the USA is the “tendency to privatize the benefits
and socialize the costs”. These are situations such as federally financed irrigation, where the
benefits are concentrated in the hands of a few farmers and the costs are widely dispersed
over society. Such behavior would be in the upper right hand box in Table 1, and one would
expect Client Politics to prevail. If so, this may be an acceptable situation, since regimes which
distribute benefits and costs widely lead to Majoritarian Politics, and hence, to inertia and
under-investment. There are obviously large overlaps among the ETP, the Polis model, and the
interest group model in their analysis of any particular situation.

6.7 Governance failures


An underlying theme of social science literature is that all governing structures fail and all
markets and hierarchies have their limitations and also fail. More effective governance regimes
or systems need to be designed/created to overcome government failure, market failure and
Water governance, water security and water sustainability 27

Table 2. Framing regimes through public choice theory.

Distribution of Distribution of costs of state intervention


benefits of state
intervention Concentrated Diffused

Concentrated Interest group politics Client politics


Organized lobby activity: high Organized lobby activity: high
but contradictory. but one-sided.
Expected outcome: Expected outcome: stable
deadlock, compromise, capture.
policy see-saw.

Diffused Entrepreneurial Politics Majoritarian Politics


Organized lobby activity: low Organized lobby activity: low.
unless policy entrepreneur
intervenes.
Expected outcome: Inertia Expected outcome: Inertia
bias, may be offset by bias except after calamity.
entrepreneur activity.

Source: Judith Rees (pers. comm., 2001).

system failure or a combination of these. For example, water is not a simple economic good;
it is sometimes a public good, sometimes a private good and often lies somewhere in between.
Its development can lead to natural monopolies, and it presents major economic and physical
externalities, etc.
These failures are listed in Table 3. They are inherent in most countries and have to be
addressed. Institutional and communication gaps are likely to be the most difficult. An empir-
ical examination of how to overcome the problems caused by market, government and system
failures is essential for each specific setting if effective water governance is to be achieved.
There are failures that cannot be easily addressed by water sector professionals as they lay
outside the water domain: for example, national institutional structures that impede polit-
ical vision, poor mechanisms for inter-sectoral dialogue, coping with unpriced assets and
public goods such as flood control and drought management. The water community never-
theless needs to understand such external governance constraints and engage with non-water
organisations to seek solutions.

7 PRINCIPLES AND PRACTICE FOR EFFECTIVE WATER GOVERNANCE

7.1 New forms of water governance


When proposing changes to water governance systems, it is important to understand
and distinguish between the different functional levels in water management: operational,
organisational and constitutional. The first focuses on the use or control of water for spe-
cific purposes to fulfil specific needs. There are always a plethora of operational enterprises
covering municipal and industrial water supply, wastewater treatment, hydropower, irriga-
tion, environmental management, tourism, etc., and they can be in public or private hands.
28 P. Rogers

Table 3. Sources of market, government, and system failure.

Sources of market failure:


Externalities (environmental, economic, and social).
Unpriced assets and missing markets.
Public goods.
Economies-of-scale.
Transaction costs.
Property rights.
Ignorance and uncertainty.
Short-sightedness.
Irreversibility.
Existence of monopolistic situations.
Sources of government failure:
Failure to correct market distortions.
Price regulation.
Subsidies to resource users and polluters.
Inappropriate tax incentives and credits.
Over-regulation or under-regulation.
Bureaucratic obstacles or inertia.
Conflicting regulatory regimes.
Short-sightedness.
Voter ignorance and imperfect information.
Special-interest effect.
Little entrepreneurial incentive for internal efficiency.
Imprecise reflection of consumer preferences.
The inability of the government to control and regulate sustainable use.
The lack of payment to other social and environmental services linked
to water.
The independence and impartiality of the regulatory agencies.
The lack of effective knowledge of the resource, the demands imposed
on the resource, and the current uses that are made of it.
Sources of system failure:
Gaps in the institutional structure that impedes the positive use of
politics.
Absence of mechanisms for coordination, decision, and conflict
resolution.
Lack of effective mechanisms for intersectoral dialogue.
Lack of mechanism for participation of the community and interested
parties.

The organisational level co-ordinates and reduces conflict between these competing enter-
prises, administers the rules and polices for water use and the users in a water system. This
function resides within the public sector – and includes for example river basin authorities
and regulatory bodies – the latter should be autonomous (within constitutional boundaries) if
they are to act impartially. Finally, the constitutional function creates the enabling environ-
ment within which the other functions operate. It sets the policies and legislation, taking into
account external governance and political imperatives. In many countries such functions are
unclear and often governments may be unable or unwilling to exercise their responsibilities. In
this case ad hoc arrangements at local government or community level are often established.
Water governance, water security and water sustainability 29

These are vulnerable as they may lack any formal basis and can be adversely affected by
vested interests or by central government policies and laws. A participatory and consultative
approach when reforming water governance systems can help to strengthen local govern-
ment and bring the positive aspects of such arrangements into the formal system and reduce
vulnerability.
Hydro-geographical boundaries – the river basin – often provide opportunities for modern
governance networks. A basin is a closed region where there are incentives for people to come
to an agreement on governance systems with water as the focus. Although basins cut across
formal jurisdictional boundaries and thus local government and other government entities that
do not necessarily work together, the basin society (a river basin agency or commission) could
require them to do so. The basin society may thus have specific governing capacities and needs.
National governments acting alone cannot easily allocate and regulate water in a basin, as they
are unlikely to appreciate local interests or priorities. Government should, however, provide the
rules and regulations and establish a framework for local people to meet (for example, the basin
community has a spatial footprint such as in the Catchment Management Agencies in South
Africa and the River Basin Agencies in France). Regulation within a basin must address issues
of quality as well as allocate quantity to users. Regulation of sectoral users such as agriculture
and industry is very weak. Preventing pollution from agricultural water use (salinity, nitrates
in groundwater) and from industries such as tanneries and mining is becoming increasingly
important, and Pakistan has recently recognised the need to regulate irrigated agriculture.
Catchment planning and management, combining land and water use, is a means to regulate
at the basin level but hitherto the tools have not been readily available to make this practical.
New approaches, for example as in the European Union (EU) Water Framework Directive and
the Streamflow Reduction Strategies in South Africa, are now starting to incorporate this into
governance systems.

7.2 Lower water use, lower conflict levels


Demand for water can be reduced voluntarily by using many different technical, social, and
economic tools. Essentially, this means that the consumer will change his or her consumption
preferences. Regulatory instruments involving permits, restrictions, and allocations to various
users and uses can also reduce water demand.
It is obvious that the water crises are due to an increase in demand and reducing that
demand would help greatly even though there would still be problems of existing levels of
resource conflicts and environmental degradation. For example, total water demand in the
USA has declined from a high in 1980, despite large increases in wealth and population
(Hutson et al., 2004). This means that maintaining aquatic environmental quality is getting
progressively easier. In this case direct water pricing policies have not brought about this
decline. It appears to be largely due to external factors such as higher energy costs and mandated
energy efficiency improvements to domestic and commercial water appliances and decline in
the value of irrigated crops. Specific water policy measures such as effluent limitations on
wastewater discharges and enforcement of federal in-stream water requirements for ecosystem
maintenance have also had a significant impact. It is worth noting how well-informed public
pressure acted as a driver for policy change and technological innovation to achieve water
savings. Each person reduced his or her water use, and overall, this has made a big difference
in water availability in the USA.
30 P. Rogers

7.3 Governance external to the water sector


Water governance can draw strength from the governance structures obtaining in other sectors
in the country, for example through the stabilisation of property rights, broad rules and laws.
Certain more general Californian state laws for example aided the creation of Californian
groundwater basins. The end of apartheid in South Africa facilitated significant changes to
water laws, and the accession of Eastern European countries to the European Union has acted
as a spur to improved water governance. Conversely, if the service provider succeeds, it can
also validate and strengthen the politics that made it possible. There are several examples of
water governance influencing external governance. The best known of these is perhaps the
co-operative water development in the Netherlands in the early part of the 20th century, which
was an important part of nation building for the modern Dutch welfare state.
Water governance traditionally begins from the social and economic policies set by gov-
ernment. However, with the growing liberalisation of trade, water services are becoming
increasingly affected by international trade agreements. Often Trade Ministry officials who
know little about water and may not necessarily consult with water officials negotiate such
trade agreements. Recent concern has been expressed by some NGOs about the inclusion of
water services in the General Agreement on Trade and Services (GATS). Whilst liberalisation
of such services may be beneficial in raising foreign direct investment, countries need to take
care in negotiating the rules under the GATS. Government negotiators can place limitations
on the commitments it makes in a specific service sector thus restricting the application of
GATS rules but this is a complex issue and often developing country negotiators are in a weak
position in such negotiations.
Of particular concern is the conflict between promoting trade and protecting the regulatory
rights of national government. It is accepted by all that the ability of government to regulate
water services providers is essential for effective private or public sector provision of water
services but the government’s right to regulate may be restricted under GATS.Apart from GATS
other trade agreements such as North American Free Trade Agreement (NAFTA), can affect
water. For example, the negotiations recently started on the Doha Round of talks on agricultural
trade liberalisation could impact on water use for food production. Similarly, debt repayments
and HIPC (Heavily Indebted Poor Countries) agreements may skew a government’s ability to
allocate budgetary provisions for water services.
Approaches to water governance should be:
• Open and Transparent: Institutions should work in an open manner. They should use lan-
guage that is accessible and understandable for the general public to increase confidence in
complex institutions.
• Inclusive and Communicative: The quality, relevance and effectiveness of government pol-
icies depend on ensuring wide participation throughout the policy chain – from conception
to implementation.
• Coherent and Integrative: Policies and action must be coherent. The need for harmony and
coherence in governance is increasing as the range of tasks has grown and become more
diverse.
• Equitable and Ethical: All men and women should have opportunities to improve or maintain
their well-being. Equity between and among the various interest groups, stakeholders,
and consumer-voters needs to be carefully monitored throughout the process of policy
development and implementation.
Water governance, water security and water sustainability 31

Performance and operation of water governance should be:


• Accountable: Roles in the legislative and executive processes need to be clear. Each of the
institutions must explain and take responsibility for what it does. The rules of the game need
to be clearly spelled out, as should the consequences for violation of the rules, and have
built-in arbitration enforcing mechanisms to ensure that satisfactory solutions can still be
reached when seemingly irreconcilable conflicts arise among the stakeholders.
• Efficient: Classical economic theory demands efficiency in terms of economic efficiency,
but there are also concepts of political, social, and environmental efficiency which need to
be balanced against simple economic efficiency.
• Responsive and Sustainable: Policies must deliver what is needed on the basis of demand,
clear objectives, an evaluation of future impact, and, where available, of past experience.
Water governance must serve future as well as present users of water services.

8 CHALLENGING THE WATER CRISIS

At any time in the history of the development of resources management there is the suspicion
that, somehow or other, things are not working out as originally planned. At the outset of the
21st century this is true of many areas such as energy, agriculture, and climate, and is partic-
ularly true for water resources management. It is tempting to say that everything that could be
said about water management has already been said. A couple of centuries of diligent research
and development have provided scientific concepts, technology, laws, and management insti-
tutions which, by and large, have served to meet most of society’s material needs. So, why
challenge this wisdom, and why now? It turns out, however, that the application of conventional
approaches has been shown to be ineffective in protecting the resource even for the current
situation and to be widely viewed as not being able to address future needs and demands, under
a highly uncertain future. The ever increasing need to spend large sums to correct mistakes of
excess pollution, or inappropriate allocation of the resource, made in the recent past tells us
that even under the best conditions (in the developed countries) our approaches are lagging
behind needs, and under the worst conditions (the developing countries) have produced major
problems which cannot be easily solved. All of these concerns are magnified by the suspicion
that we may be entering a period of rapid human-induced climate change with very uncer-
tain implications for water resource management. Despite what may occur in the future, the
overhang of existing problems in the developed and developing countries seems to necessitate
fairly radical reevaluation of the prevailing approaches that has placed us in this situation.

8.1 In defense of conventional wisdom


In spite of the indictment of conventional wisdom above, I believe that it has within its purview
all the tools needed to help humankind survive the water resources challenges posed. It is often
that the conventional wisdom, if properly examined, already contains the seeds for challenging
the way the water community currently does business. I believe that conventional wisdom and
conventionally accepted concepts and ideas about water could, if understood in a holistic way,
significantly improve the sustainability of water as a resource and as a pillar of the ecosystem
within which we all live in the face of the real possibilities of a water crisis. Part of the problem
is that conventional wisdom is not so widely understood as it appears and not so widely applied;
in other words, it is not conventionally accepted conventional wisdom!
32 P. Rogers

8.2 Nine important facts of life for water


The nine very important facts-of-life concerning water discussed above, are convention-
ally well known, but generally not understood by all of the different water constituencies.
They are certainly conventional wisdom, but never seem to be quoted together as the fabric
or the framework under which to view water problems. These important facts cover a mul-
titude of issues, but by themselves they constitute a sort of Nine Commandments of water
sustainability!
– Blue/Green/Brown water: It is estimated that human appropriations of the green water were
almost three times as large as those of the blue water. The possibility of using more of the
green water and reducing the demands on the blue water are active questions that could lead
to dealing more easily with feeding the future populations.
– Asymmetries in water use: For most uses small amounts of water are actually consumed
with the bulk returning, albeit in many cases polluted, to the water source. Irrigation users,
however, typically evaporate more than 70% of the withdrawn water. Industry, particularly
heavy industry and fossil fuel electric generation, withdraw huge amounts of water, but con-
sume little. Municipal and commercial withdrawals are typically the smallest withdrawals
with about 20% consumptive use. This leads to the role of irrigation water as a balancing
mechanism for meeting water demands in other sectors. A 10% reduction in water use by
irrigation would more than double the amounts needed for municipal and industrial uses.
– Irrigation flywheel: In many countries of the world as much as 60–95% of all water is
consumed by irrigated agriculture. This fact is extremely important when looking toward
future potential water shortages. The question then hinges on whether or not such a 10%
reduction in irrigation water use could be achieved by efficiency improvements alone, or
the equivalent areas for rainfed crops and additional trade need to be developed, in order to
meet increasing demands for food, fiber, and other industrial crops.
– Virtual water escape hatch: Directly related to the irrigation flywheel is what Allan (2001)
characterized as virtual water: the use of imports of food crops as a substitute for use of
domestic water for irrigation. Many arid countries in the world already rely upon the trade
in virtual water to meet their needs for food. The question remains: as to what extent can
the trade be expanded by developing the use of more green water?
– Low cost desalination breakthrough: Recent developments in membrane filtration tech-
nologies combined with reverse osmosis have revolutionized the potential for widespread
application of desalination. While this does not give an instant solution to the problem
in many poor cities in the developing world, it does indicate that reasonable economically
achievable technology is available to solve the urban water supply problems as they develop.
– The ecological revolution: The major reason for widespread dissatisfaction about water
issues in developing countries stems from the vast numbers who have no access to adequate
drinking water or sanitation. It is simply beyond the financial capabilities of existing gov-
ernmental and international agencies to cover the cost of providing conventional water and
sanitation to these persons and their increasing numbers. Recent developments and cost
breakthroughs in ecological sanitation practices that use little or no water, point the way to
resolving this major issue.
– Transboundary conflicts: Since there is no strong international law governing the resolution
of transboundary water quality and quantity disputes, the potential for conflict remains
high. This issue needs careful consideration by riparians and the regional family of nations to
Water governance, water security and water sustainability 33

head off these conflicts from becoming worse. New approaches through regional governance
organizations such as the EU, SAARC (South Asian Association for Regional Cooperation),
NAFTA, SADAC (Southern African Development Community), seem to hold the best
chance of resolving the issues.
– Uncertainty of water availability: By the end of the 20th century, hydrologists were search-
ing for evidence of long-term anthropogenic climate change caused by greenhouse gases.
Many conflicting data sets and theories have been reviewed with contradictory results and
while at present there is no general agreement as to the change in variability of precipi-
tation over the next 5 decades, there seems to be a growing consensus that precipitation
will increase over large areas of the globe and that sea level may rise between 0.1 and 0.3
meters. Most of the discussion about the uncertain future of water has focused upon the
water supply side of the equation; much larger uncertainties can be expected on the water
demand side influenced by demographic parameters, economic growth, technology change,
and life styles. There is a pressing need for water planners and agencies to improve their
approaches to resolving resource issues with such high levels of uncertainty.
– Idiosyncrasy of water institutions: Finally, one water fact-of-life makes it very difficult to
implement the types of technical and social breakthroughs already available; the hurdle
is that property rights, institutions, and laws, vary widely from country to country, even
within countries. The legal and political institutions tend to get in the way of rational water
management. In many cases it is impossible to reallocate or trade between conflicting water
uses. The confusion allows for the entrenchment of bad conventional institutional views
and blocks consideration of newer approaches.

9 CONCLUSIONS

This chapter has provided an introduction to the issues of water security and scarcity in the
context of governance of the resource. The discussion points to many issues which are not yet
resolved in analyzing the question of crisis or reality? It does, however, make a strong plea
for the consideration of the political and social dimensions of the question. Later chapters in
the book deal with the physical, technical, and economic aspects of the water crisis, but the
reader is urged not to ignore the cautionary parts of this first chapter. The water crisis is real if
the players involved in development and management of water ignore the powerful influence
of politics on how the problems are structured and that the outcomes depend upon the political
structuring of the problem.
The commentator on this chapter at the workshop in Santander (Domingo Jiménez Beltrán),
summarized the main points of the chapter as follows:

• Sustainability is now a well-established concept that contains both the more restricted con-
cepts of security and survivability, and makes redundant any approaches that are solely
aimed at them.
• By establishing sustainable development you can improve both the environment and the
resource availability and at a lower cost.
• Sustainability and governance is one and the same thing. If the principles of good gov-
ernance (efficacy, efficiency, equity, coherence, transparency, accountability, and public
participation) are pursued, then sustainability must be a consequence.
34 P. Rogers

• There is no water crisis; there is a clear water management crisis. This is demonstrated by
the failure to bring water use onto a sustainable path.
• The principles and practices for future water governance are well known and tested, but it
seems to be difficult to escape from the weight of past practices. New forms of governance
(Nueva Cultura del Agua in Spain) using a different logic from the ones that created the
current situation are desparately needed.
• Most, if not all, of the needed technical, economic, political, and social tools needed to
effect this transition are already well known. What is needed is some way to synthesize
all of these components into a coherent whole. This should be the role of Integrated Water
Resources Management (IWRM).

REFERENCES

Alcamo, J.; Heinrichs, T. & Rösch, T. (2000). World Water in 2025. World Water Series, Report 2. Centre
for Environmental Systems Research, University of Kassel, Germany.
Allan, J.A. (2001). The Middle East Water Question: Hydropolitics and the Global Economy. I.B. Tauris,
London, UK.
Bornstein, D. (2004). How to Change the World: Social Entrepreneurs and the Power of New Ideas.
Oxford University Press.
Buchanan, J. & Tullock, G. (1962). Calculus of Consent: Logical Foundations of a Constitutional
Democracy. University of Michigan Press, Michigan, USA.
Crenson, M.A. & Ginsberg, B. (2002). Downsizing Democracy: How America Sidelined its Citizens and
Privatized its Public. Johns Hopkins.
Dahl, R.A. (1961). Who Governs. Yale University Press.
Diamond, J. (2005). Collapse: how societies choose to fail or succeed. Viking, New York, USA.
Dorfman, R.; Jacoby, H.D. & Thomas, H.A. Jr. (eds.) (1972). Models for Managing Regional Water
Quality. Harvard Press.
Downs, A. (1957). An Economic Theory of Democracy. Harper and Row.
Falkenmark, M. & Rockstrom, J. (2004). Balancing Water for Humans and Nature. EARTHSCAN,
London, UK.
Falkenmark, M. & Widerstrand, C. (1992). Population and Water Resources: A delicate Balance.
Population Bulletin, 47(3). Population Reference Bureau, Washington, D.C., USA.
GWP (Global Water Partnership) (2000). Towards Water Security: A Framework for Action. GWP,
Stockholm, Sweden.
Hoekstra, A.Y. & Hung, P.Q. (2002). Virtual Water Trade, a quantification of virtual water flows between
nations in relation to international crop trade. Technical Report. Virtual Water Research Report Series,
No. 11. IHE, Delft, the Netherlands.
Hutson, S.S.; Barber, N.L.; Kenny, J.F.; Linsey, K.S.; Lumia, D.S. & Maupin, M.A. (2004).
Estimated Use of Water in the United States in 2000. USGS (United States Geological Survey),
Circular 1268.
IPCC (Intergovernmental Panel on Climate Change) (2001). Climate Change 2001: Third Assessment
Report. Cambridge University Press.
IWMI (2000). World Water Supply and Demand. International Water Management Institute, Colombo,
Sri Lanka.
Kaufmann, D.; Kraay, A. & Zoido-Lobatón, P. (1999). Governance Matters. Policy Research Working
Paper 2196. World Bank Institute, October, 64 pp.
Lasswell, H.D. (1936). Politics: Who gets What, When and How? McGraw-Hill.
L’vovich, M.I. (1979). World Water Resources and their Future. Translation by the American Geophysical
Union. LithoCrafters Inc., Chealsea, Michigan, USA.
Maass, A. (1951). Muddy Waters: The Army Engineers and the Nations Rivers. Da Capo Press.
Maass, A. & Anderson, R.L. (1978). … and the desert shall rejoice: conflict, growth, and justice in arid
environments. MIT Press.
Water governance, water security and water sustainability 35

Maass, A.; Hufschmidt, M.M.; Dorfman, R.; Thomas. H.A. Jr.; Marglin, S.A. & Fair, G.M. (1962).
Design of Water Resources Systems: New Techniques for Relating Economic Objectives and
Engineering. Harvard University Press.
Olson, M. (1965). The Logic of Collective Action: Public Goods and the Theory of Groups. Harvard
University Press.
Olters, J.P. (2001). Modeling Politics with Economic Tools: A Critical Survey of the Literature. Working
Paper 01/00. International Monetary Fund.
Ostrom, E. (1990). Governing the Commons: The Evolution of Institutions for Collective Action.
Cambridge University Press.
Pezzey, J.C.V. (1992). Sustainability Development Concepts: An Economic Analysis. World Bank,
Environment Paper no. 2. Washington, D.C., USA.
Pezzey, J.C.V. (2002). Sustainability Policy and Environmental Policy. Economics and Environment
Network, Australian National University. Working Paper, EEN0211. Canberra, Australia: 43 pp.
Pierre, J. (ed.) (2000). Debating Governance, Authority, Steering and Democracy. Oxford University
Press.
Population Action International (2003). Sustaining Water: Population and the Future of Renewable Water
Supplies. Population Action International, Washington, D.C., USA.
Postel, S.L.; Daily, G.C. & Erhlich, P.R. (1996). Human Appropriation of Renewable Fresh Water.
Science, 271: 785–788.
Putnam, R.D. (2000). Bowling Alone: The Collapse and Revival of American Community. Simon and
Schuster.
Ramirez-Vallejo, J. & Rogers, P. (2004). Virtual Water Flows and Trade Liberalization. Journal of Water
Science and Technology, 49(7): 25–32.
Repetto, R. (ed.) (1986). World Enough and Time. Yale University Press.
Rogers, P. & Hall, A. (2003). Effective Water Governance. Global Water Partnership. TEC Background
Paper No. 7. Stockholm, Sweden.
Shafik, N. & Bandyopadhyay, S. (1992). Economic Growth and Environmental Quality: Time-Series
and Cross-Country Evidence. Background paper for the World Development Report. WPS 904. World
Bank, June.
Shiklomanov, I.A. (2000). Appraisal and Assessment of World Water Resources. Water International,
25(1): 11–32.
Shiklomanov, I.A. & Sokolov, A.A. (1983). Methodological basis of world water balance investigation
and computation. In: New Approaches in Water Balance Computations. Proceedings of the Hamburg
Symposium. International Association for Hydrological Sciences. Publication No. 148.
Solow, R.M. (1991). Sustainability: An Economist’s Perspective. The eighteenth J. Seward Johnson
Lecture. Woods Hole Oceanographic Institution, Massachusetts, USA.
Stone, D.A. (2000). Policy Paradox: The Art of Political Decision Making. W.W. Norton & Company,
New York, USA.
Tendler, J. (1997). Good Governance in the Tropics. Johns Hopkins Press.
UN/WWAP (United Nations/World Water Assessment Programme) (2003). Water for People, Water for
Life. UN World Water Development Report. UNESCO (United Nations Educational, Scientific and
Cultural Organization) and Berghahn Books.
U.S. Water Resources Council (1983). Principles and Guidelines. Executive Order 11747, February 3.
Wade, R. (1987). The Management of Common Property Resources: Finding a Cooperative Solution.
World Bank Research Observer, 2(2): 219–234.
Waterbury, J. (1979). Hydropolitics of the Nile Valley. Syracuse University Press, Syracuse, New York,
USA.
WMO (World Meteorological Organization) (1992). The Dublin Principles Keynote Papers. International
Conference on Water and Environment: Development Issues for the 21st Century (Dublin, Ireland).
WMO, Geneva, Switzerland.
Wolf, A.; Shira, T.; Yoffe, B. & Giordano, M. (2003). International waters: identifying basins at risk.
Water Policy, 5(1): 29–60.
World Commission on Environment and Development (1987). Our Common Future. Oxford University
Press, New York, USA.
World Resources Institute (2003). Water Resources and Freshwater Ecosystems. Earth Trends (World
Resources Institute environmental information portal).
CHAPTER 2

Water for growth and security

W.J. Cosgrove
Bureau d’Audiences Publiques sur l’Environnement, Québec, Canada

ABSTRACT: The lack of action to address the need for better management of the world’s scarce
water resources, and in particular about the lack of investment in water services and infrastructure that
are essential to economic and social development should be of deep concern to all. We could say that the
efforts of the water community in raising awareness of the critical state of the world’s water resources and
of the need for water for health and development have succeeded. Yet the reality is that picture remains
bleak. Progress can be made with an approach having three basic elements: (1) Clear goals that target
the poor and use the right means; (2) Partnerships; and (3) Using processes underway.

Keywords: Water resources, infrastructure gap, economic development, targeting poor, benefits

1 INTRODUCTION

I am deeply concerned about the lack of action to address the need for better management
of the world’s scarce water resources, and in particular about the lack of investment in water
services and infrastructure that are essential to economic and social development. I have for
many years seen the connection between economic development and the presence of water
infrastructure.

2 TO-DAY’S SITUATION

On the planet to-day over 1000 million live in absolute poverty, with incomes of less than
one US$/day. Thousands of millions of people lack access to safe water supply and sanitation.
Rural communities lack access to water for subsistence farming. Droughts and floods wreak
havoc and destroy years of hard work and savings. Scarce resources are polluted and overused.
The natural sources of environmental services and of life to the species with which we share
this planet are being destroyed.
The international community is not unaware of this situation. For example, the Millennium
General Assembly of the United Nations (UN) adopted a number of Millennium Development
Goals (MDGs). Among them was a goal to reduce by 2015 half the proportion of people without
sustainable access to safe drinking water. The World Summit on Sustainable Development
(WSSD) in Johannesburg two years later added a goal to reduce by half the population without
access to basic sanitation by 2015. Leaders gathered at WSSD also called upon all countries
to prepare country plans for Integrated Water Resource Management (IWRM) by 2005. In
Kyoto in 2003 at the 9th Meeting of the Committee of the Parties (COP9) those responsible for
38 W.J. Cosgrove

Figure 1. Cause-effect chains and inter-linkages between water and the MDGs.

implementing the Kyoto Protocol recognized that adaptation to climate variability and change
is an emerging issue for water managers.
The UN Task Force on Water for the Millennium Development Goals has recognized the
links between water and achievement of nearly all of the MDGs (Figure 1: UN Task Force on
Water and Sanitation, 2004).
Thus there is no doubt that water has been recognized as a critical factor on the Sustainable
Development agenda. We could say that the efforts of the water community in raising awareness
of the critical state of the world’s water resources and of the need for water for health and
development have succeeded.
Yet the reality is that the bleak picture remains that I described in the opening of this
chapter and there are few signs that it is changing. Former President Jimmy Carter of the USA
in his speech accepting the Nobel Peace Prize (Carter, 2002) said:
“Citizens of the 10 wealthiest countries are now 75 times richer than those who live in the 10
poorest ones. The separation is increasing every year. Not only between nations, but within
Water for growth and security 39

them. The results of this disparity are the root causes of most of the world’s unresolved
problems, including starvation, illiteracy, environmental degradation, violent conflict, and
unnecessary illnesses that range from guinea worm to HIV and AIDS”.

He might well have mentioned lack of access to water for life and livelihoods.

3 ADDRESSING THE GAP

Water management (supply, demand and availability) is a universal concern and response
requires local and global cooperation in partnerships between governments at all levels, civil
society, private sector and knowledge institutes. Progress can be made with an approach with
three basic elements:
– Clear Goals
• Targeting the poor
• Using the right means
– Partnerships
– Using processes underway.

3.1 Clear goals


3.1.1 Targeting the poor
The UN Convention states that special regard should be given to the requirements of vital
human needs. In this context world community thinks of Africa. In Africa only 175 million
out of a rural population of 410 million have access to an improved water supply. The total
number to be served will have increased through population growth to 517 million by 2015.
The situation in urban areas appears to be better. Reports show that 175 million out of a total
population of 215 million currently have access to an improved source. However through
population increase and rural-urban migration, the number requiring service by 2015 will
have increased to 389 million. Moreover, the definition used of those with/without service is
technological. Those in urban areas connected to a distribution system are considered to be
served. Yet recent reports demonstrate that in many cases these distribution systems are not
operated reliably, with no water in them much of the time and the water that is there often
polluted. The figures for access to basic sanitation services and hygienic living conditions are
much worse. The absolute numbers of those without access are higher in Asia. However it is
the appalling high percentages that are seen in Africa.

3.1.2 Targeting means: the infrastructure gap


Statistics show that as a level of country’s income grows the value of its infrastructure increases.
For example Chad which in 1990 had a per capita income (1985 PPP1 terms) of roughly
US$ 200, had invested only US$ 10 per capita in infrastructure through its history. Australia,
with a per capita income about 600 times as high in the same year had invested close to
US$ 8000 per capita in infrastructure. The poor countries of Africa are grouped together with
low investments in infrastructure. The industrialised countries with high incomes have invested

1
Purchasing power parity.
40 W.J. Cosgrove

large amounts in infrastructure. But have these countries invested more as their economies
grew, or did the countries’ economies grow when they invested in infrastructure? Experience
in many places has shown that it is the latter. Consider, for example, the Tennessee Valley
in the USA where the investments in infrastructure by the Tennessee Valley Authority clearly
lifted the residents of the valley out of poverty as well as contributing to development beyond
the valley through power generation.
I share the hypothesis put forward by David Grey, Senior Water Adviser of the World Bank,
that water security is essential for sustained economic growth and poverty eradication and that
there is a minimum platform of water resources infrastructure and institutions to achieve water
security. Developed countries have invested public funds heavily to provide this infrastructure
that is a public good.

3.1.3 Needs and benefits


This infrastructure stores water to reduce flooding and provide sources of supply during dry
seasons. The need for such infrastructure is especially great in Africa. In Kenya, for example,
fluctuations in rainfall may vary plus and minus 40% year on year. Subsequent flood and
drought years between 1998 and 2000 have been estimated to have resulted in losses to the
Kenyan economy of 22% of the average Gross Domestic Product (GDP) of the country over
those two years. A correlation between rainfall and GDP has been clearly demonstrated in
Zimbabwe, and should be expected in any country in which agriculture counts for a significant
percentage of total GDP.
The provision of water supply and sanitations services is perhaps a greater priority given
the reduction in premature deaths, sickness and other hardships and indignities that will be
eliminated, and the benefits from a healthier population at study and at work. Often overlooked
is the fact that investment in infrastructure and services is not just an expense. Enterprises,
whether public or private, are necessary to produce the materials required, to construct the
systems and to operate and maintain them. Their establishment and strengthening add to a
country’s institutional framework and benefit further development.

3.2 Partnerships
Partnerships can provide a major means of accelerating investment in water services and
infrastructure. These may be at the national, regional (continental) or international level.
Many governments have begun the process of decentralizing responsibility for the con-
struction and management of services to local governments, including water and sanitation
services. This approach has great potential for increasing the rate at which improved access
to these services is achieved and maintained. However, it requires that the capacity be created
at the local level, and it further requires that mechanisms be found to bring the financial
resources to this level. The international and multinational banks are looking at ways to do
this through sub-sovereign lending. However I believe that the solutions to these questions
will best be found within countries, with national governments searching for appropriate solu-
tions in partnership with local elected officials. The local governments, in turn, are working
more and more in partnership with community-based organizations and non-governmental
organizations.
Moreover, the solution is not to be found in the traditional approach that requires that the
capacity be created first before investments are made. Rather investments must be forthcoming
Water for growth and security 41

that are in keeping with available capacity but stretch it, so that capacity to build and manage
services grows as the size of the systems grow.
The most prominent example at the regional level is without a doubt the New Partnership
for Africa’s Development (NEPAD). African heads of state have committed themselves to a
participatory development approach and to being held accountable for their progress by their
peers. African governments and civil society are partners in this process. African water min-
isters are working together in the African Ministerial Conference on Water (AMCOW). They
are supported by UN Water for Africa and encouraged by other members of an independent
group known as the African Water Task Force. The African Development Bank has responded
to requests from NEPAD and AMCOW by the establishment of the African Water Facility to
accelerate investments and innovation in the water sector, and the Rural Water Supply and
Sanitation Initiative. Like-minded donors and investors are committed to responding with
assistance to this commitment by Africans to help themselves.
Internationally a major development was the creation of Type II Partnerships at the World
Summit on Sustainable Development in Johannesburg. One of these, tentatively titled the
North-North Partnership, aims at developing water partnerships within developed countries,
linking these to share their practices and to make links to countries of the South. Many of
the latter have already developed national water partnerships under the aegis of the Global
Water Partnership. A similar partnership has been at work through the Cooperative Program on
Water and Climate (formerly the Dialogue on Water and Climate). This network brings together
climatologists, meteorologists, hydrologists and water managers from national and basin level
from both developed and developing countries to address issues of water management in the
face of climate variability and long-term climate change.

3.3 Using processes underway


Water management (supply, demand and availability) is a universal concern. As described
above successful action requires local and global cooperation in partnerships between gov-
ernments at all levels, civil society, private sector and knowledge institutes. Progress can be
made using processes underway if the goals are clear, appropriate means are employed and all
work in partnership.
Many governments have not yet included investments in water among their priorities. One
interpretation of this can be that they have not yet understood the importance of water to
achieving all of their economic and social development goals. We know that awareness of the
needs and benefits are well understood by ministers responsible for the sector. However we
must all work at impressing on the ministers of finance and economic development the critical
role that water plays in all sectors. Clear targets for access to water services must be supported
by short- and long-term budgets, policies, and investment strategies shared by all members of
government.
At the intergovernmental level one of the important existing processes is that of the Commis-
sion on Sustainable Development (CSD). The current cycle of the CSD deals with water supply
and sanitation and housing. The meeting of CSD 13 in April 2005 provides an opportunity for
countries to demonstrate that they have already begun to fulfill their responsibilities using the
national processes described above. CSD 13 should not be a place where ministers of the envir-
onment from around the world put forward proposals for policy approaches to be followed
by developing countries in managing their water, but rather as an occasion for developing
42 W.J. Cosgrove

countries to put forward the policies that they know will work and to ask the international
community to support them in their implementation.
The United Nations General Assembly in September 2005 marking the fifth anniversary of
the adoption of the Millennium Development Goals (MDGs) provides another such occasion.
However, it will be even more important as it will permit the heads of state of developing
countries to refer to the policies that they have presented at CSD 13 and to ask their partners
of the developed world to fulfill theirs.
In March 2006 the World Water Council will sponsor the Fourth World Water Forum
hosted by the Government of Mexico in Mexico City. Its overall theme will be Local Actions
to Address Global Challenges. This reflects the recognition that it is local actions that unleash
local social energy. The Forum will support the institutional networks that support local actors
and processes, hopefully developing synergies and increasing collaboration between different
stakeholder groups. National and local governments and their partners in civil society will
have an opportunity to present cases that demonstrate how their new policy approaches are
being translated into actions on the ground.

REFERENCES

Carter, J. (2002). The Nobel Peace Prize Lecture. December 10, Oslo, Norway.
United Nations Task Force on Water and Sanitation (2004). Achieving the Millennium Develop-
ment Goals for Water and Sanitation: What Will It Take? Interim Full Report. Task Force on
Water and Sanitation. Millennium Project. February. [www.unmillenniumproject.org/documents/
tf 7interim.pdf].
CHAPTER 3

Collective systems for water management: is the Tragedy


of the Commons a myth?

E. Schlager
School of Public Administration and Policy, University of Arizona, Tucson, USA

E. López-Gunn
London School of Economics and Political Science, UK

ABSTRACT: The tragedy of the commons is a myth in water management. Hundreds of case studies
of local water users devising institutional arrangements to successfully govern their use of shared water
resources have been documented. Failures in water governance that have occurred are rarely due to the
tragedy of the commons, that is, to the lack of institutional arrangements. Failures are more often due to
the challenges of devising, adapting, and maintaining institutional arrangements in dynamic water set-
tings. This chapter explores two institutional challenges that repeatedly confront water users – devising
institutional arrangements that are well matched to the physical and social features of the water setting;
and devising complementary and supportive relations among organizations and governments operating
at different scales. The chapter begins with a review of the literature on common pool resources and
the creation and maintenance of local, self-governing institutional arrangements illustrated through case
studies of water governance. It then turns to an examination of cross-scale linkages – relations among
organizations and governments at different scales, from the local to the international. The chapter pro-
vides an in-depth examination of cross-scale linkages, what appear to be the most important features
of such linkages, and, in particular, important features of higher level governments that promote sup-
port for the self-governing efforts of local resource users. Cases of water governance from the USA
and Spain are used to illustrate cross-scale linkages and the role they play in common pool resource
management.

Keywords: Institutions, self-governance, cross-scale linkages, common pool resources, multi-level


governance, water governance

1 INTRODUCTION

The model of the tragedy of the commons predicts the failure of appropriators to sustainably use
a common pool resource. Appropriators race to harvest what they can from a resource, helpless
to prevent the race from occurring. Appropriators are trapped in a Hobbesian state of nature,
incapable of communicating and, consequently, unable to devise agreements or institutional
arrangements that would allow them to sustainably govern a common pool resource.
While all commons tragedies result in failures, not all failures are due to commons tragedies.
In devising long enduring, self-governing arrangements for managing water, appropriators
must not only pay close attention to carefully matching institutional arrangements with the
physical and social setting, but also to the interactions among different levels of organization
44 E. Schlager, E. López-Gunn

and government. In other words, multiple institutional hurdles must be safely traversed for
long enduring, self-governing arrangements to emerge and to be sustained.
Failures in water governance are rarely due to the tragedy of the commons, that is, to the
lack of institutional arrangements. Failures are more often due to the challenges of devising,
adapting, and maintaining institutional arrangements in dynamic water settings. Increasingly
in this context, sustainability is perceived as a dynamically maintained equilibrium, not a
static one (Agrawal, 2002). This chapter explores two institutional challenges that repeatedly
confront water appropriators – devising institutional arrangements that are well matched to the
physical and social features of the water setting; and devising complementary and supportive
relations among organizations and governments operating at different scales.
Well-matched institutional arrangements require that local water users, with their wealth
of local time and place information, participate in the creation, maintenance, and revision of
water institutions. For example, questions over whether institutions are created top down or
bottom up can have implications at the local scale on the ownership and empowerment local
people might feel towards these institutions. Local level participation is necessary, but often
not sufficient, for sustainable governance.
The spatial extent of water resources often means that water is not just a local affair, indeed in
some cases it is a highly conflictive transboundary international issue. Many local commu-
nities, regional governments, and national governments participate in water governance. In a
globalized world, institutions are increasingly nested within one another and “cross-scale insti-
tutions and vertical linkages become even more important” (Berkes, 2002: 294). Issues like
whether higher level authorities crowd in or, on the contrary, facilitate better local governance
of water resources is progressively more important, i.e. whether interdependencies are sym-
metrical or benign or unidirectional (and malign) (Young, 2002). The quality of the cross-scale
linkages among organizations and governments directly affects the ability of water appropria-
tors to govern and use water resources in a sustainable fashion. As Lowndes & Wilson (2003:
279) state: “partnerships and networks are as important as hierarchical intra-organizational
relationships”.
In the following two sections the literature on common pool resources and the creation
and maintenance of local, self-governing institutional arrangements is surveyed and illus-
trated through case studies of water governance. These two sections provide the foundation
for the heart of the paper – cross-scale linkages. Cross-scale linkages refer to relations among
organizations and governments at different scales, from the local to the international. Now
that considerable progress has been made in recognizing and explaining local, self-governing
efforts, increasing attention is being devoted to the context in which local institutional arrange-
ments operate, or cross-scale linkages. The last section of the chapter provides an in-depth
examination of cross-scale linkages, what appear to be the most important features of such
linkages, and, in particular, important features of higher level governments that promote sup-
port for the self-governing efforts of local resource users. Cases of water governance from the
USA and Spain are used to illustrate cross-scale linkages and the role they play in common
pool resource management.

2 ROBUST LOCAL SELF-GOVERNING INSTITUTIONAL ARRANGEMENTS

As the title of this chapter implies, for decades the most widely accepted explanation for natural
resource degradation and destruction was Garrett Hardin’s (1968) model of the tragedy of
Collective systems for water management: is the Tragedy of the Commons a myth? 45

the commons. It dovetailed nicely with other widely accepted models of individual decision-
making in the social sciences. For instance, Olson’s (1965) theory of collective action predicted
that only under very special circumstances would people be able to overcome free riding
problems and engage in collective action. From game theory, the prisoners’ dilemma game
made equally grim predictions: people will pursue their own self-interest, even if in doing so
they make themselves worse off than if they had cooperated and pursued the group’s interest.
Natural resource economics models also highlighted conflicts between individual self-interest
and the efficient use of natural resources (Gordon, 1954; Scott, 1955; Anderson, 1986). All
models pointed in a single direction – self-interested individuals will make choices and take
actions that will lead to the economic ruin, if not the actual destruction, of natural resources.
This simplistic view of individual rationality as self-interested is now much more nuanced,
as “homo reciprocans, agents are neither completely self-interested not blatantly altruis-
tic … agents as homo reciprocans in the sense that people typically exhibit altruistic tendencies
only towards those who cooperate with them and reciprocate their altruistic gestures”
(Onyeiwu & Jones, 2003: 236). This means that while in some cases grim predictions were
borne out, in others cases quite the opposite occurred. Users of shared resources cooperated
to define and administer self-governing institutional arrangements that allowed them to coor-
dinate their actions and use resources in sustainable ways (McCay & Acheson, 1987; Berkes,
1989). Individuals are complex, and the challenge is to increase the numbers of conditional
cooperators (Ostrom, 2000a) through strong institutional arrangements.
Based on in-depth and carefully constructed analyses of numerous cases of both successful
and failed efforts at local, self-governance of natural resources, Ostrom (1990) posited a set
of institutional design principles that appeared to be common among all cases of successful,
long-lasting, efforts at local self-governance (see Table 1). Design principle one, exclusion,
is critical if appropriators are to commit to following a set of institutional arrangements over
time and investing in modifying them as circumstances warrant. Exclusion, while critical, is
insufficient to ensure long-term commitment to the rules. The rules themselves must make
sense, they must be crafted to the exigencies of the situation, and as the situation changes, the
appropriators must have the ability to modify the rules. Accountable monitors and graduated

Table 1. Ostrom’s Institutional Design Principles for long enduring Common Pool Resource (CPR)
arrangements (Ostrom, 1990: 90).

1 Individuals or households who have rights to withdraw resource units from the CPR must be
clearly defined, as must the boundaries of the CPR itself.
2 Appropriation rules restricting time, place, technology, and/or quantity of resource units are related
to local conditions and to provision rules requiring labor, material, and/or money.
3 Most individuals affected by the operational rules can participate in modifying the operational rules.
4 Monitors, who actively audit CPR conditions and appropriator behavior, are accountable to the
appropriators or are the appropriators.
5 Appropriators who violate operational rules are likely to be assessed graduated sanctions
(depending on the seriousness and context of the offense) by other appropriators, by officials
accountable to these appropriators, or by both.
6 Appropriators and their officials have rapid access to low-cost local arenas to resolve conflicts
among appropriators or between appropriators and officials.
7 The rights of appropriators to devise their own institutions are not challenged by external
governmental authorities.
8 Appropriation, provision, monitoring, enforcement, conflict resolution, and governance
activities are organized in multiple layers of nested enterprises.
46 E. Schlager, E. López-Gunn

sanctioning maintain appropriators’ commitment to institutional arrangements. Knowing that


most appropriators are following the rules most of the time sustains rule following behavior.
In instances of rule violations, graduated sanctions act to bring appropriators’ actions in line
with the rules. Finally, conflict resolution mechanisms and at least a minimal recognition of
the right to organize prevent these institutional arrangements from unraveling due to internal
strife or invasion from external governmental authorities.
The principles have received considerable support from in-depth research conducted across
several different types of common pool resources. In relation to water, studies of irrigation
systems and groundwater basins both support the conclusion that resource users can devise and
manage robust self-governing arrangements. Irrigation studies have confirmed the importance
of exclusion and carefully crafted rules (design principles one and two), and the importance of
appropriators participating in devising and modifying rules (design principle three) (Tang,
1989, 1992, 1994; Lam, 1998). Tang studied 43 irrigation systems, examining the dif-
ferences between high performing systems, those that were well-maintained and in which
farmers followed the water allocation rules, and low performing systems, those that either
were not well maintained and/or farmers failed to follow the rules consistently. Tang (1994)
argues that among irrigation systems that perform well, rules that govern water allocation and
maintenance activities are better crafted to the specific conditions of each irrigation system.
High performing systems were associated with multiple rules that adequately limited access
to the system and that fairly allocated water among the irrigators. Poorly performing irrigation
systems were characterized by a single simple rule set or by no rules at all. Access to the irriga-
tion systems was not adequately regulated and water allocation rules often did not work well.
While low performing irrigation systems were similar in the rules of access and allocation
that were used, such systems were much more likely to be government owned rather than
farmer owned. Government officials, since they are not directly subject to the irrigation rules
that they devise, face few incentives to design rules that ensure the effective operation of
irrigation systems. Instead, they face incentives to devise rules that increase their political
support and that lighten their administrative burdens. The rationalist institutionalist approach
takes a dim view of civil servants acting as self-interested individuals. Conversely, because
farmers directly experience the consequences of their rule-making decisions, they confront
incentives to carefully craft the rules to the particular situation that they face (Tang, 1992).
Tang’s research was replicated and extended by Lam (1998). Extensive data on 150 Nepalese
irrigation systems was analyzed by Lam (Lam et al., 1994; Lam, 1998). Like Tang (1992,
1994), Lam (1998) found that farmer managed irrigation systems performed significantly
better than did government managed irrigation systems. Irrigators in farmer managed systems
exhibited significantly higher levels of: (1) entrepreneurial activities in attempting to coor-
dinate irrigation activities; (2) information and understanding of the irrigation system; and
(3) mutual trust (Lam, 1998: 26–133). Irrigators in farmer managed systems also used more
varied and complex sets of rules for governing their activities.
Equally in Spain in the context of groundwater management, a classic example of a common
pool resource, water users associations in the same region have had very different experiences
in mutually beneficial collective action. The water users associations that were created top
down (in Western Mancha and Campo de Montiel) have not fared well compared to the
one created bottom up (in Eastern Mancha). In Eastern Mancha where farmers themselves
have designed and created their own institutions, with the support of higher-level authorities,
empowerment has led to farmers slowly devising their own rules. Thus in these three cases,
Collective systems for water management: is the Tragedy of the Commons a myth? 47

Table 2. Comparative application of Ostrom (1992) design principles in three spanish aquifers.

Western
Mancha Campo de Eastern
Design principles (Ostrom, 1992) aquifer Montiel Mancha

1. Clearly defined boundaries. Both 

Geographic
the boundaries of the service area 
 (aquifer boundary)   
and the individuals or household with
rights to use water from an irrigation 


system area clearly defined. Water rights × × ×
2. Collective choice agreements. Most individuals × × ×
affected by operational rules are included in the group
that can modify these rules.
3. Monitoring. Monitors, who actively audit physical × × 
conditions and irrigator behavior, are accountable to the
users and/or are the users themselves.
4. Graduated sanctions. Users who violate operational × × 
rules are likely to receive graduated sanctions.
5. Conflict resolution mechanisms. Users and their officials × × 
have rapid access to low-cost local arenas to resolve
conflict between users or between users and officials.
6. Minimal recognition of rights to organize. The rights   
of users to devise their own institutions are challenged
by external government authorities.
7. Nested Enterprises. Appropriation, provision, monitoring, × × 
enforcement, conflict resolution, and governance activities
are organized in multiple layers of nested enterprises.

the design principles identified by Ostrom (1992) have been a good predictor of the success
or failure of collective action and self-governance (see Table 2). In the absence of these
design principles (i.e. farmer self-management) higher level authorities have attempted to
fill the gap of farmer initiatives through perverse subsidies to change behavior, effectively
crowding out farmers’ initiatives, and generating and re-enforcing a rent-seeking behavior
which then creates a path dependency (Pierson, 2000) difficult to change (López-Gunn, 2003a,
2003b).
The empirical evidence is overwhelming. Resource users are capable of devising rules
that carefully guide and constrain their harvesting activities in sustainable ways. Resource
users are not inevitably trapped in a tragedy of the commons. Furthermore, the most resilient
institutional arrangements devised by local level resource users are those with well-defined
boundaries with resource users as the primary decision-makers. Such arrangements are well
matched to the problems and challenges that the resource users face and they function in a
manner that supports continued interaction and cooperation. Monitoring supports commitment
to the rules by providing assurances that most users are following the rules most of the time.
When rule violations are detected, graduated sanctions are used as a means of correcting the
actions of the rule breaker and not as a means of punishment. Also, there is a recognition that
conflict and disagreements are inevitable and that some means of peaceably and legitimately
settling differences is necessary.
48 E. Schlager, E. López-Gunn

It is important to stress that although research and analysis of internal collective arrange-
ments are plentiful, the relevance of context and external factors impacting on these internal
arrangements was taken as given or not explored in depth. Yet due to the nature of complex
networks which now dominate institutional arrangements, the focus is now shifting towards
a deeper understanding of the two design principles identified by Ostrom (1990), centered on
recognition of the right to organize and nested enterprises. For example, the acknowledgement
by higher-level authorities of the right of appropriators to devise their own institutional arrange-
ments is in itself an indicator of a very basic (but crucial) trust cementing cross-scale linkages.
In the context of theories on institutional learning, some basic level of social capital that can
be built upon, harnesses local knowledge and creativity (Lowndes & Wilson, 2003: 287).

3 THE EMERGENCE OF LOCAL SELF-GOVERNING ARRANGEMENTS

Identifying critical features of long enduring self-governing institutional arrangements differs


from identifying the conditions under which resource users may be willing to invest in develop-
ing or revising such arrangements. For institutional arrangements to persist and to perform well
over time, they must be revised to address new challenges and changing circumstances. After
considerable additional research, Ostrom (2000b) has suggested a set of conditions that support
the emergence of cooperation to devise or revise institutional arrangements. These conditions
consist of attributes of common pool resources and attributes of appropriators (seeTable 3). The
attributes of common pool resources focus on the ability of appropriators to readily understand
the resource and their actions in relation to it. For instance, indicators, predictability and spatial
extent are supportive of resource users learning the critical features and underlying dynamics
of a resource. In this line there are recent developments on communities developing their own
set of community indicators to help with local water management particularly grounded on
that community. Such information is vital if resource users are to develop rules well matched to
the physical setting. Feasibility of improvement, however, is the critical attribute. Whether it is
worthwhile investing in knowledge of the resource and in institutional arrangements governing
the resource depends on the returns to be realized. If changes in behavior will have little effect
because the resource is too degraded, resource users are unlikely to take such steps.
Even if resource users know and understand the resource well, they are only likely to
engage in institutional change, according to Ostrom (2000b) if they exhibit certain attributes.
If they are heavily dependent on the resource and if they envision themselves, their children,
grandchildren and community remaining heavily dependent on the resource well into the
future, then they are more likely to engage in rule development, particularly if they share a
common understanding of the resource. Conversely, if they are only moderately dependent
on the resource for income, or if they envision their children pursuing alternative income
generating activities, then they may be less willing to invest in new rules. Furthermore, the
resource users must trust one another and they must have leaders or organizational skills to draw
upon to engage in rule development. If they do not trust one another, if they perceive that other
resource users will not consistently follow the rules, or will act in ways that undermine coopera-
tive relationships, then they are unlikely to invest in institutional arrangements. Engaging in
institutional change is no easy matter. Resource users must be committed to the resource, they
must have a relatively well-developed understanding of it, they must trust one another, and
they must have sufficient leadership and autonomy to engage in the task of rule change.
Collective systems for water management: is the Tragedy of the Commons a myth? 49

Table 3. Ostrom’s attributes that support institutional change (Ostrom, 2000b: 40).

Attributes of the resource


1 Feasible improvement: Resource conditions are not at a point of deterioration such at it is useless
to organize or so underutilized that little advantage results from organizing.
2 Indicators: Reliable and valid indicators of the condition of the resource system are frequently
available at a relatively low cost.
3 Predictability: The flow of resource units is relatively predictable.
4 Spatial extent: The resource system is sufficiently small, given the transportation and
communication technology in use, that appropriators can develop accurate knowledge of
external boundaries and internal microenvironments.
Attributes of appropriators
1 Salience: Appropriators are dependent on the resource system for a major portion of their
livelihood or other important activity.
2 Common understanding: Appropriators have a shared image of how the resource system
operates … and how their actions affect each other and the resource system.
3 Low discount rate: Appropriators use a sufficiently low discount rate in relation to future
benefits to be achieved from the resource.
4 Trust and reciprocity: Appropriators trust one another to keep promises and relate to one
another with reciprocity.
5 Autonomy: Appropriators are able to determine access and harvesting rules without external
authorities countermanding them.
6 Prior organizational experience and local leadership: Appropriators have learned at least
minimal skills of organization and leadership through participation in other local associations
or studying ways that neighboring groups have organized.

Whether these attributes of the resource and of appropriators are key to explaining insti-
tutional change remains to be determined. Increasingly, however, there is more and more
evidence of the crucial role social capital (both bonding and bridging) can play in cementing
different institutional arrangements to foster collaborative behavior in water management. As
Onyeiwu & Jones (2003: 243) state: “social capital appears to be a built in mechanism, albeit
a subtle one for ensuring cooperative behavior”. The same authors identify a series of fea-
tures in social capital: (1) fostering mutually beneficial exchanges through the development
of norms of civic behavior, trust and cooperation; (2) establishing a consultative and collec-
tive decision-making process that reduces externalities and promotes the reduction of public
goods; and (3) reducing the transaction costs that also facilitate the flow of information and
innovation. That is, cooperative behavior is underpinned by social capital, and there is a direct
proportional relationship between social capital (both bonding and bridging) and collaborative
behavior to encourage rational egoists to behave like (conditional) cooperators.
While considerable research has investigated robust institutional arrangements, little
research has focused on institutional change and stability in relation to common pool resources.
Research is needed on questions related to institutional learning like studies centered on triple
loop learning (Maarveland & Dangbegnon, 1999). In single loop learning the outcomes of
decision-making are evaluated in the context of goals and expectations, in double loop learn-
ing, feedback starts to generate change in current practices, however, in triple loop learning
institutions learn to learn, or institutional reflexiveness. This is compared to its opposite,
learned helplessness when individuals or communities are unable to influence their contexts
through their behavior. Once again in institutional learning cross-scale linkages are crucial to
foster triple loop learning and avoid learned helplessness.
50 E. Schlager, E. López-Gunn

The example of common pool resource management demonstrates why it is crucial that
water users themselves are involved in management leaving room for experimentation and
trial and error. If allowed opportunities to experiment, water users can progress to double
loop learning and the development of adaptive management. As Maarveland & Dangbeg-
non (1999) state: “space needs to be created to diversify learning patterns”. In this context,
the key relevance of higher-level authorities’ involvement to strengthen or rejuvenate local
management becomes apparent. The role of higher level authorities cannot and should not
be underestimated; in providing support, granting legitimacy to local institutions, creating
enabling legislation and backing capacity building (Berkes, 2002). A delicate balance needs
to be established between push and pull, to create space for social learning in common pool
resource management.

4 NEXT STEPS IN COMMON POOL RESOURCE RESEARCH

The most critical tasks presently facing researchers are theory development and testing
(Agrawal, 2002; Stern et al., 2002). Agrawal argues that too much attention is still being
devoted to the identification of factors that support self-governance and sustainable use of
common pool resources, so much so that researchers face an almost embarrassing wealth
of factors. Agrawal (2002: 62–63) has identified 33 distinct factors that a variety of scholars
argue account for success. So many possible factors present a variety of problems. First, lists of
factors alone do not explain success or failure in common pool resource management. Rather,
the factors are configural, they interact, and through that interaction outcomes are realized.
To date, however, little attention has been devoted to identifying or exploring the interactions
among the factors – that is, to use the factors as a means to develop theory. Second, the
factors rarely apply uniformly across all common pool resource settings. The values that the
factors take and their effects on an outcome are influenced by the larger context within which
the common pool resource is situated. The factors are contingent on a variety of contextual
variables, such as ties to regional, national and global markets, or the linkages and relations
among different levels of government. A key factor that keeps coming to the fore is social
capital cementing cross-scale institutional arrangements, leading to the sustainability of the
commons – the durability of institutions that frame the governance of common pool resources
(Agrawal, 2002: 44).

4.1 Cross-scale linkages


One of the more important tasks facing water policy researchers is to develop theoret-
ically grounded explanations of the relations among organizations and governments that use,
manage, and govern water. The spatial extent of water resources often means that water is not
just a local affair. Many local communities, regional governments, and national governments
participate in water governance. The quality of the cross-scale linkages among organizations
and governments directly affects the ability of water appropriators to govern and use water
resources in a sustainable fashion.
The importance of cross-scale linkages is demonstrated in research by Blomquist et al.,
(2004) that examined and compared the conjunctive water management activities of local
jurisdictions among three USA states – California, Arizona, and Colorado. Conjunctive water
Collective systems for water management: is the Tragedy of the Commons a myth? 51

management refers to the coordinated use of surface water and groundwater as a means of
developing additional water supplies and limiting the waste of existing supplies. Typically, the
coordination of the two water sources occurs through recharge projects. Surplus surface water
is stored, or recharged, in an underground basin for use at a later time.
Local water appropriators contemplating investing in recharge projects must have assur-
ances that the water they store underground will be available to them at a later time. In other
words, appropriators must be able to exclude those who did not contribute to their project from
accessing their stored water. The spatial extent of a basin must be such that water appropriators
can gain control over the basin and exercise exclusion if they are to cooperate to engage in
conjunctive water management.
In California, conjunctive water management activities occur only in basins in which appro-
priators have organized themselves and have developed a set of institutional arrangements
whereby they control access to and use of the basins (Heikkila, 2001; Blomquist et al.,
2004). Surprisingly, however, even though appropriators in Arizona and Colorado do not
control access to and use of groundwater basins, they too actively engage in conjunctive water
management. Why? The larger institutional setting – specifically cross-scale linkages. The
governments of Arizona and Colorado, unlike the government of California, have devised and
allocated private property rights in surface water, groundwater, and in stored, or recharged,
groundwater. Appropriators in Arizona and Colorado, because of their state granted private
property rights in water, are assured that if they store water underground they will be able
to retrieve it at a later date because of their property rights in that water. They do not first
have to gain control over a groundwater basin. Appropriators in California, on the other hand,
must first gain control over their basin and define property rights in it before they are suf-
ficiently secure to invest in conjunctive water management projects. Thus, the necessity of
appropriators controlling access to a common pool resource depends on the larger institutional
setting–context matters.
Research in Spain highlights that enduring institutional arrangements are dependent on the
positive synergy between the regional government, water authorities and the water users them-
selves. Debates that center on property rights and the trade off between flexibility and security,
are mirrored on current debates on institutional arrangements and the trade off between robust-
ness and flexibility. The comparison of three Spanish aquifers mentioned above and their
institutional arrangements to solve aquifer over-use are an interesting example. In the case
of Western Mancha, institutional arrangements are in a state of flux and therefore are flex-
ible, however, institutional arrangements are not robust, which is translated into free-riding,
corruption and aquifer over-use. By contrast, the institutional arrangements in the neigh-
boring aquifer, Campo de Montiel, are robust, yet not flexible enough to acknowledge the
legitimate concerns of non-farming water uses. The third case is more promising because the
institutional arrangements being developed are robust, yet also flexible. Why the difference?
The answer lies in a case of earned autonomy on the part of the farmers, as a successful story
of co-management of water resources, and thus rides on the strength of cross-scale linkages.
The concept of earned autonomy refers to the performance of local government and how – if
it delivers in terms of policy outcomes – higher level authorities will grant further autonomy.
Greater attention to cross-scale linkages potentially signals a major change in the social
sciences. As Lowndes & Wilson (2003) comment: “the postmodern paradox is based on
the idea that institutional fragmentation and multilevel interdependencies is now the norm in
governance. In this context eclecticism is an institutional design principle whose time has come
52 E. Schlager, E. López-Gunn

and which aims at encouraging reflexivity and learning and signals the end of modernist grand
narratives about the right way to govern”. Or as Rhodes (1997: xii) states: “messy problems
require messy solutions and there is no right institutional fix, it’s the mix that matters in the
end governing code”.
Young (2002) argues that greater attention must be paid to the quality and complementarity
of the relationships among organizations. According to Young (2002), the issue is not so much
getting the tasks divided up appropriately among different levels of government; rather the issue
is more one of developing supportive and cooperative relations, or cross-scale linkages, among
organizations and governments at different levels of society. The quality of relations among
organizations and governments appears to depend on a variety of factors such as features of an
organization, authority relations among organizations, and interactions among organizations.

4.1.1 Features of organizations


Young (2002) argues that the ability of an organization, such as a national government, to make
commitments to other governments and to follow through with those commitments depends
on three different features. Capacity refers to material resources and social capital necessary
to realize the activities the government has committed itself to (Young, 2002: 278). If the
government does not possess the resources or does not allocate the resources to meet its com-
mitment then, according to Young, it is lacking in capacity. Compatibility refers to the degree
of congruence between the commitments made by a government and the internal structure and
functioning of the government and society (Young, 2002: 278). If a government that oversees
a centrally planned economy agrees to engage in market activities, the compatibility between
how it operates and how it has committed to act is relatively low. Finally, competence refers to
the legal and political authority needed to make commitments to others (Young, 2002: 277).
As Young notes, competence depends on the structure of the government. The competence of
federal systems is relatively low because agreement must be gained from different branches
of government or from sub-national governments. Young (2002) focuses on the characteristics
of organizations as critically defining cross-scale linkages. Capacity, compatibility, and com-
petence together determine the ability of an organization to make and follow through with its
commitments to other organizations. However, Young’s notion of competence suggests that
authority relations among organizations also play an important role.

4.1.2 Authority relations


The ability of a government to follow through with its commitments depends on its relations
with sub-governments and with citizens. Conversely, the ability of citizens to engage in gov-
erning common pool resources depends on the authority and autonomy granted them by a
government. If resources users have constitutionally protected property rights in a resource
they are more able to protect their interests than if they do not have such protections. Their
protected right to be in the resource provides them with considerable leverage in determining
who else can and cannot be in the resource. Furthermore, the extent to which higher levels
of government can influence the collective choice processes of local appropriators depends
on the decision-making authority that local appropriators have been granted and the extent to
which that authority is protected. If the authority is embedded in legislation or in a constitu-
tion, higher levels of government will be more limited in their ability to directly affect local
level choices. In order to override local decisions, external government officials will have to
change legislation and/or change constitutional provisions.
Collective systems for water management: is the Tragedy of the Commons a myth? 53

The quality of cross-scale linkages is also determined by features of higher-level govern-


ments. Local, self-governing efforts are supported by higher-level governments that provide
technical expertise, funding, and independent and dispassionate monitoring and conflict
resolution mechanisms (Ostrom, 1990; Blomquist, 1992; Landre & Travis, 1998). Local self-
governing efforts are much more likely to be successful if appropriators have considerable
autonomy to devise their own governing arrangements, secure from arbitrary interventions
from higher levels of government and if higher levels of government provide supportive
contexts that encourage self-governing activity.
Institutional features that support local level self-governing activity, however, present real
challenges for higher-level governments engaged in addressing larger scale common pool
resource problems that local communities alone cannot readily address. Higher-level govern-
ments will find it much more difficult to command compliance from relatively autonomous
local resource users. Rather than commanding compliance, higher-level governments may
have to rely more heavily on other mechanisms to bring the activities of local resource users
in line with their interests and goals. Enabling legislation, grants, funding for projects are
mechanisms that higher-level governments can use to try to align local interests with their own.

4.1.3 Interactions among organizations


Scharpf (1997) suggests that relations and interactions are likely to differ considerably among
organizations depending on whether they are attempting to coordinate their actions to reach
a better collective outcome or whether they are locked in a zero-sum game. Relations among
organizations are much more likely to be complementary and cooperative if they are attempting
to realize a shared benefit than if they view one another as strict competitors. Even if actors
realize that they need to work together to create a shared benefit, relations may be tense
because of what Scharpf labels the negotiators’ dilemma. In a situation in which all actors’
contributions are important for realizing a desired outcome, actors often must address two
closely related issues: production – what the final project or outcome will be, and distribution –
how the benefits and costs will be allocated among the participants. Conflict may be intense
as the parties address production and distribution issues, or the creation and sharing of value
(Scharpf, 1997: 120). According to Scharpf, conflict is likely to be manageable if organizations
are bargaining over the production of a desired outcome and not over the distribution of the
costs and benefits or realizing the outcome. If, however, organizations are bargaining over
the distribution of benefits and costs, conflict may become so intense that agreements are not
reached and cooperation and coordination do not occur.
The literature on cross-scale linkages is relatively sparse. Scholars who have attended to
such issues suggest that attention be paid to features of the organizations, authority relations
among them, and the dynamics of their interaction in order to explain the role of cross-scale
linkages in supporting or hindering local level self-governance.

4.1.4 Case studies: attributes of higher-level authorities


According to Young (2002: 266) “the extent to which environmental or resource regimes
yield outcomes that are sustainable is a function not only of the allocation of tasks between
and among institutions operating at different levels of social organizations, but also of cross-
scale interactions among distinct institutional arrangements … it seldom makes sense to focus
exclusively on finding the right level or scale at which to address specific problems … the key
to success lies in allocating specific tasks to the appropriate level”.
54 E. Schlager, E. López-Gunn

This section will argue that a key to sustainability lies in strengthening cross-scale linkages,
and in particular, paying careful attention to encouraging certain characteristics in higher level
authorities. In Spain, the key role of higher level authorities can be seen in the three neighboring
aquifers already discussed, with quite distinct and complex internal institutional arrangements
yet very similar higher level authorities. This chapter will not analyze the internal factors
described by Ostrom (1990, 1992), which are described in a previous paper (López-Gunn,
2003a). Rather the analysis focuses on the key role played by external, higher level authorities
in strengthening or weakening cross-scale linkages.
The case study concentrates on three aquifers located in the Mancha region of Spain: the
Western Mancha aquifer, the Eastern Mancha aquifer, and the Campo de Montiel aquifer.
The three aquifers are relatively large and are heavily utilized mainly for irrigation, which
boomed from the 1970s leading to cases of severe aquifer drawdown. However, the insti-
tutional response has been very different both internally in terms of the activities of their
respective water user associations and externally in relation to their interaction with higher
level authorities and cross-linkages with state and non-state actors. The three case studies
exemplify the complex set of networks that now characterizes water (Rogers & Hall, 2003).
This section analyses attributes of higher level authorities that can favor or hinder mutually
beneficial collective action to prevent or halt aquifer over-use. It identifies what is called
attributes of higher level authorities, which are listed below:

(1) Clear boundaries in water rights: It is essential that as far as practicable higher-level
authorities facilitate the clear definition of property rights in water. Ideally, this should be
undertaken in joint collaboration with water users, who have the benefit of local knowl-
edge. In the Spanish case studies the situation in relation to clear definition of water
rights (and therefore boundaries) is very different. In the case of Western Mancha the
situation is chaotic with estimated thousands of illegal abstractors. By contrast, in the
case of both the Eastern Mancha and the Campo de Montiel slowly boundaries are being
established for property rights. The reason for this success compared to the Western Man-
cha by and large lies in the collaboration between farmers and the water authorities to
determine these water rights. However, it is in the case of the Eastern Mancha where
boundaries are actually being drawn jointly in a clear case of a partnership drawing on the
synergy between localized knowledge and regulatory power. By contrast, in the case of
the Campo de Montiel, although boundaries are clear, the foundations of the relationship
between higher level authorities and farmers is not necessarily very positive and bound-
aries have been defined because farmers themselves saw it in their self-interest to do so in
order to exclude other users. Therefore, this brings us to the next attribute of higher level
authorities.
(2) Legitimization of appropriators: Ostrom (1992) already identified as a factor the key
importance of the recognition of the right to organize. Local organization cannot operate
in a vacuum, and in the long term it is essential that legitimization is granted to local
organizations from higher-level authorities. Local organization can be legitimate yet have
no supports from higher-level authorities. This in the longer term will harm the prospects
of these institutions. In the case of the Campo de Montiel and the Western Mancha water
user associations have not been acknowledged fully by the water authority and in both
cases, despite the strong internal institutional arrangements of these water user associ-
ations, this has led to conflict, which has undermined the activities of these water user
Collective systems for water management: is the Tragedy of the Commons a myth? 55

associations, either by leading to litigation or through time spent over intractable water
issues.
(3) Facilitates and support initiatives of appropriators: The example above has to be con-
trasted with the experience in the Eastern Mancha, where both the regional government
and the water authority have been very supportive and facilitated initiatives from the water
user association leading, in the first instance to partnership agreements, at a later stage to
co-management and eventually hopefully to self-governance.
(4) Builds trust with other cross-scale linkages: Yet the question remains why have in some
cases higher level authorities, like Eastern Mancha facilitated and acknowledged the legit-
imacy of water users and supported local institutions, whilst in other cases, particularly
in the Western Mancha it has been the opposite (a latent conflict that spills over from
time to time). It is argued that the key difference, as stated in previous sections in the
chapter has been bridging trust cementing cross-scale linkages. It is important to stress
the dynamic nature of social capital, i.e. it can be eroded and it can be created. In the
Eastern Mancha, it is clear that third party trust was created between the water user asso-
ciations and higher-level authorities and a key ingredient was the leadership role played
by individuals at different scales.
(5) Clear division of responsibilities: Equally, it is increasingly important to provide, as
far as possible, the subsidiarity principle, decisions should be taken at the appropri-
ate level. However, echoing Young (2002), it does not necessarily mean just finding
the right scale but rather clarity in the interplay of increasing functional dependencies.
For example, in many occasions, more so in the context of common pool resources, it
might be relevant to “combine the strength of government level and local level resource
management, to mitigate the weaknesses sin of each” (Berkes, 2002: 301). For exam-
ple, favoring from the outset a co-management approach as a partnership between local
users and agencies focuses not only on outcomes but also on the process itself as social
learning.
(6) If relevant; from supervisory control, to co-management to self-regulation ( from extrinsic
reward to intrinsic motivation): Increasingly in the policy making arena where the shift
has left government behind, to be overtaken by governance it is more and more pertinent
for higher level authorities to be more explicit (and upfront?) on asymmetrical power rela-
tions and institutional bargaining. For example, acknowledge more directly issues related
to access, use, management, exclusion and transferability of water resources, and the rules
that might (or might not) be up for negotiation. A networked society calls for a new type
of institution imbued with a particular culture.
(7) Institutional culture: The right institutional culture when faced with the increasing pattern
of decentralization prevailing in much of the developing world in natural resource manage-
ment (including water) would “seek to harness rather than override the local knowledge
and creativity of a multiplicity of designers” (Lowndes & Wilson, 2003: 287). Institu-
tional culture should as far as possible, be receptive, flexible and robust. For instance, the
dominant institutional culture in higher level authorities should support positive patterns
of behavior (such as community leadership) and eradicate negative traditions (such as
departmentalism, paternalism, and social exclusion) (Lowndes & Wilson, 2003). Most
recent research in common pool resource management pinpoints to a renaissance in
local level governance and new pushes for decentralization. This chapter attempts to
contextualize this renaissance, since local institutions do not operate in a vacuum, and
56 E. Schlager, E. López-Gunn

in particular to cross-scale linkages and the role of higher level authorities and external
events.

5 CONCLUSION

The challenge facing the tragedy of the water commons is to design robust yet flexible insti-
tutional arrangements. This chapter has argued that increasingly the focus of attention is
shifting to questions on the strength of cross-scale linkages. This is not surprising, since the
21st century the questions posed by water turn around water governance not government
and ever more complex set of networks surrounding mutually beneficial collective water
management.
In the context of institutional cross-scale linkages, the challenge lies in developing cross-
scale linkages that are capable of delivering adaptive management – uncertainty is integrated
into the decision-making process and institutions that are capable of learning from successes
and failures.
For cross-scale linkages it might be safer to opt for a more post-modernist view of policy, or
of historical institutionalism (as compared to rational institutionalism) where institutions are
less abstracted and more contextualized; i.e. “where actions are not the product of calculated
decisions, rather they are embedded in institutional structures of rules, norms, expectations
and traditions that severely constrain the behavior of social actors” (Mule, 1999: 148). Strong
institutions and strong cross-scale linkages that can cope with critical junctures and unantic-
ipated consequences (Eaton, 2004). In this context, more research should be undertaken on
questions centered on institutional change and stability, and by this one has to distinguish,
as Lowndes & Wilson (2003: 280) state: “between organizational change and institutional
change, while the former may involve no more than structural re-organization, the latter
requires effective rules of behavior are altered through specifying and embedding new norms,
incentives and sanctions”. In the coming decades one will need to assume change and therefore
stability will have to be explained (Berkes, 2002).
However, questions still remain on external and endogenous changes from critical junctures,
and on the speed of change. For example, local institutions as Berkes (2002) comments are
more likely to adapt over a period of decades, rather than months, yet higher-level authorities
increasingly operate on election cycles.
In a globalized world, and almost as a counter-reaction, attention is turning towards local
level institutions rooted in local circumstance and heavily contextualized. It is important to
distinguish local governance from local government and that “top down and bottom up institu-
tional influences interact in important ways. The extent of local distinctiveness in governance
arrangements is related to the degree of autonomy and diversity that higher levels of govern-
ment will tolerate” (Lowndes & Wilson, 2003: 280). Equally one has to distinguish between
decentralization of political authority and decentralization of governing capacity. In this second
case it is a question of institutional change. In more and more countries of the world, local
level common property institutions are the new institutional messiah; local institutions can be
robust, flexible, reflexive, varied; they can be quick to learn, innovate and adapt compared to
centralized institutions (Berkes, 2002). However, this chapter has argued that the Achilles heel
of these institutions – potentially crucial for integrated water management – are cross-scale
linkages and the role of higher level authorities.
Collective systems for water management: is the Tragedy of the Commons a myth? 57

REFERENCES

Agrawal, A. (2002). Common Resources and Institutional Sustainability. In: E. Ostrom, T. Dietz,
N. Dolsak, P.C. Stern, S. Stonich & E.U. Weber (eds.), The Drama of the Commons. National Academy
Press, Washington, D.C., USA: 41–86.
Anderson, L. (1986). The Economics of Fisheries Management. Rev. ed. Baltimore, Maryland, USA.
Johns Hopkins University Press.
Berkes, F. (ed.) (1989). Common Property Resources: Ecology and Community-based Sustainable
Development. Belhaven, London, UK.
Berkes, F. (2002). Cross-Scale Institutional Linkages: Perspectives from the Bottom Up. In: E. Ostrom,
T. Dietz, N. Dolsak, P.C. Stern, S. Stonich & E.U. Weber (eds.), The Drama of the Commons. National
Academy Press, Washington, D.C., USA: 293–322.
Blomquist, W. (1992). Dividing the Waters: Governing Groundwater in Southern California. ICS Press.
San Francisco, California, USA.
Blomquist, W.; Schlager, E. & Heikkila, T. (2004). Common Waters, Diverging Streams. Resources for
the Future. Washington, D.C., USA.
Eaton, K. (2004). Designing sub-national institutions: Regional and municipal reforms in post
authoritarian Chile. Comparative political Studies, 37(2): 218–244.
Gordon, H.S. (1954). The Economic Theory of a Common Property Resource: the Fishery. Journal of
Political Economy, 62: 124–142.
Hardin, G. (1968). The Tragedy of the Commons. Science, 162: 1243–1248.
Heikkila, T. (2001). Managing Common Pool Resources in a Public Service Industry: The Case of
Conjunctive Water Management. Ph.D. Dissertation, University of Arizona, USA.
Lam, W.F. (1998). Governing Irrigation Systems in Nepal: Institutions, Infrastructure and Collective
Action. ICS Press. San Francisco, California, USA.
Lam, W.F.; Lee, M. & Ostrom, E. (1994). An Institutional Analysis Approach: Findings from the NIIS
on Irrigation Performance. In: J. Sowerwine, G. Shivakoti, U. Pradhan, A. Shukla & E. Ostrom (eds.),
From Farmers’ Fields to Data Fields and Back, a Synthesis of Participatory Information Systems for
Irrigation and Other Resources. International Irrigation Management Institute, Colombo, Sri Lanka;
and Institute of Agriculture and Animal Science, Rampur, Nepal: 69–93.
Landre, P. & Travis, L. (1998). Collaborative Watershed Management in the Finger Lakes Region,
New York. [http://www.indiana.edu/∼iascp/Final/landre.pdf].
López-Gunn, E. (2003a). The role of collective action in water governance: a comparative study of
groundwater user associations in La Mancha aquifers (Spain). Water International, 28(3): 367–378.
López-Gunn, E. (2003b). Policy change and learning in groundwater policy: a comparative anal-
ysis of collective action in La Mancha (Spain). Ph.D. Thesis. King’s College, University of
London, UK.
Lowndes, L. & Wilson, D. (2003). Balancing revisability and robustness? A new institutionalist
perspective on local government modernization. Public Administration, 81(2): 275–298.
Maarveland, M. & Dangbegnon, C. (1999). Managing natural resources: a social learning perspective.
Agriculture and Human Values, 16: 267–280.
McCay, B.J. & Acheson, J.M. (1987). Human Ecology of the Commons. In: B.J. McCay & J.M. Acheson
(eds.), The Question of the Commons: the Culture and Ecology of Communal Resources. University
of Arizona Press, Tucson, USA: 1–34.
Mule, R. (1999). New Institutionalism: distilling some hard core propositions in the works of Williamson
and March and Olsen. Politics, 19(3): 145–151.
Olson, M. (1965). The Logic of Collective Action: Public Goods and the Theory of Groups. Harvard
University Press, Cambridge, Massachussets, USA.
Onyeiwu, S. & Jones, R. (2003). An institutional perception of cooperative behavior. Journal of Socio-
economics, 32: 233–248.
Ostrom, E. (1990). Governing the Commons: the Evolution of Institutions for Collective Action.
Cambridge University Press, New York, USA.
Ostrom, E. (1992). Crafting institutions for self-governing irrigation systems. ICS Press, San Francisco,
California, USA.
58 E. Schlager, E. López-Gunn

Ostrom, E. (2000a). Collective action and the evolution of social norms. Journal of Economic
Perspectives, 14(3): 137–158.
Ostrom, E. (2000b). The danger of self-evident truths. PS: Political Science and Politics, 33(1): 33–44.
Pierson, P. (2000). Increasing Returns, Path Dependence, and the Study of Politics. American Political
Science Review, 94(2): 251–267.
Rhodes, R.A.W. (1997). Foreword. Managing complex networks. W. Kickert. Sage, London, UK.
Rogers, P. & Hall, P. (2003). Effective Water Governance. Global Water Partnership, Technical
Committee.
Scharpf, F. (1997). Games Real Actors Play. Westview Press, Boulder, Colorado, USA.
Scott, A. (1955). The Fishery: the objectives of sole ownership. Journal of Political Economy, 63:
116–124.
Stern, P.C.; Dietz, T.; Dolsak, N.; Ostrom, E. & Stonich, S. (2002). Knowledge and questions after 15
years of research. In: E. Ostrom, T. Dietz, N. Dolsak, P.C. Stern, S. Stonich & E.U. Weber (eds.), The
Drama of the Commons. National Academy Press, Washington, D.C., USA: 443–490.
Tang, Y.S. (1989). Institutions and Collective Action in Irrigation Systems. Ph.D. Dissertation. Indiana
University, USA.
Tang, Y.S. (1992). Institutions and Collective Action: Self-Governance in Irrigation. ICS Press, San
Francisco, California, USA.
Tang, Y.S. (1994). Institutions and Performance in Irrigation Systems. In: E. Ostrom, R. Gardner &
J. Walker (eds.), Rules, Games and Common-pool Resources. Ann Arbor, University of Michigan
Press, USA: 225–246.
Young, O.R. (2002). Institutional Interplay: the Environmental Consequences of Cross-Scale Intera-
ctions. In: E. Ostrom, T. Dietz, N. Dolsak, P.C. Stern, S. Stonich & E.U. Weber (eds.), The Drama
of the Commons. National Academy Press, Washington, D.C., USA: 263–292.
II

The Economic Value of Water


CHAPTER 4

The economic conception of water

W. M. Hanemann
University of California, Berkeley, USA

ABSTRACT: This chapter explains the economic conception of water – how economists think about
water. It consists of two main sections. First, it reviews the economic concept of value, explains how it is
measured, and discusses how this has been applied to water in various ways. Then it considers the debate
regarding whether or not water can, or should, be treated as an economic commodity, and discusses the
ways in which water is the same as, or different than, other commodities from an economic point of view.
While there are some distinctive emotive and symbolic features of water, there are also some distinctive
economic features that make the demand and supply of water different and more complex than that of
most other goods.

Keywords: Economics, value of water, water demand, water supply, water cost, pricing, allocation

1 INTRODUCTION

There is a widespread perception among water professionals today of a crisis in water resources
management. Water resources are poorly managed in many parts of the world, and many
people – especially the poor, especially those living in rural areas and in developing countries –
lack access to adequate water supply and sanitation. Moreover, this is not a new problem –
it has been recognized for a long time, yet the efforts to solve it over the past three or four
decades have been disappointing, accomplishing far less than had been expected. In addition,
in some circles there is a feeling that economics may be part of the problem. There is a sense
that economic concepts are inadequate to the task at hand, a feeling that water has value in
ways that economics fails to account for, and a concern that this could impede the formulation
of effective approaches for solving the water crisis.
My own personal assessment is that the situation is somewhat more complex than critics
suggest. On the one hand, as environmental and resource economics has evolved over the past
forty years, it has developed a conceptual toolkit that I think is well suited for dealing with many
of the issues of water supply and water resource management. On the other hand, economists
sometimes slip into older ways of thinking and characterize economic value in terms that are
inadequate or misleading. Moreover, even among economists there is an inadequate appreci-
ation of the complexities of water as an economic commodity; these render it distinctive from
other commodities and they contribute to the explanation of the current crisis in water.
This chapter examines the economic conception of water – how economists think about
water – at least partly in light of these concerns1 . It consists of two main sections. Section 2

1
I am well aware that there are other conceptions of water coming from other disciplines. I see those as
complements, rather than substitutes, for the economic conception of water. For a fascination account
of alternative conceptions of water in the 19th century, see Hamlin (2000).
62 W.M. Hanemann

reviews the economic concept of value, explains how it is measured, and discusses how this
has been applied to water in various ways. Section 3 takes on the debate regarding whether or
not water can or should be treated as an economic commodity, and discusses the ways in which
water is the same or different as other commodities. The chapter ends with a few concluding
observations in section 4.

2 WHAT IS ECONOMIC VALUE? HOW IS IT MEASURED?

Is economic value measured by market price? If an item has a price of US$ X , is this also
the amount of its economic value? Most people assume the answer is yes, and economists
sometimes also make this statement. For example, the following passage equates the economic
value of water with its market price:

“In a market system, economic values of water, defined by its price, serve as a guide to allocate
water among alternative uses, potentially directing water and its complementary resources into
uses in which they yield the greatest total economic return” (Ward & Michelsen, 2002).

If it were true that economic value is measured by market price, this would imply that only
marketed commodities can have an economic value. Items that are not sold in a market –
including the natural environment, and public goods generally – would have no economic
value. If this were so, economic value would indeed be a narrow concept and at variance with
many people’s intuitive sense of what is valuable.
In fact, however, economic value is different than price. Price does not in general measure
economic value, and items with no market price can still have a positive economic value.
This was first pointed out by Dupuit (1844) and Marshall (1879). But, as explained below, it
took until the 1970s for this to become well accepted within modern economics. It was around
this time that operational procedures became available to measure economic value separately
from price; and it was around this time that non-market valuation emerged as a field in eco-
nomics. It so happens that water as a commodity played a role in these developments, both clar-
ifying the economic concept of value and developing operational procedures for measuring it.

2.1 The meaning of economic value


The distinction between market price and economic value was famously noted by Adam Smith
in a passage in the Wealth of Nations describing the paradox of water and diamonds:
“The word value, it is to be observed, has two different meanings, and sometimes expresses
the utility of some particular object, and sometimes the power of purchasing other goods
which the possession of that object conveys. The one may be called value in use; the other,
value in exchange. The things which have the greatest value in use have frequently little or no
value in exchange; and, on the contrary, those which have the greatest value in exchange have
frequently little or no value in use. Nothing is more useful than water; but it will purchase
scarce anything; scarce anything can be had in exchange for it. A diamond, on the contrary,
has scarce any value in use; but a very great quantity of other goods may frequently be had in
exchange for it” (book I, chapter IV).

Smith was using the comparison between water and diamonds to illustrate a distinction
between two different meanings of value. In fact, neither the distinction between the definitions
The economic conception of water 63

of value nor the use of water to illustrate it was original with Smith2 . Two thousand years before
Smith, Plato had observed that: “only what is rare is valuable, and water, which is the best of
all things . . . is also the cheapest”3 . In fact, Plato and Smith were both expressing a thought
that had occurred to many other people over the ages, namely that the market price of an item
need not reflect its true value. Market price reflects the fluctuating circumstance of daily life,
whether the vagaries of supply (sudden scarcity, monopoly, etc.) or demand (temporary needs,
changes in taste, fads and fashions), while the true value is something more basic, enduring,
and stable.
Just what this true value is has been seen differently at different times. For Plato, the true
value was intrinsic to the ideal form underlying the item. For Aristotle, it was intrinsic to the
natural end that the item served. Aristotle also originated the distinction between this value –
in effect, value in use – and value in exchange: “of everything which we possess there are
two uses . . . one is the proper, and the other the improper or secondary use of it. For example,
a shoe is used for wear, and is used for exchange; both are uses of the shoe”4 . For Saint
Thomas Aquinas, the true value of an item was determined by its inner goodness, an intrinsic
quality of the item stemming from its relation to the divine purpose. In the 14th century, some
Scholastics propounded a view closer to Aristotle’s that the intrinsic value of an item arises
from its inherent usefulness and ability to please man according to rules of reason. However,
starting with Davanzati in 1588, Italian humanists stressed subjective human preference rather
than objective human need as the basis of true value. Men seek happiness, Davanzati wrote,
by satisfying all their wants and desires, and they value items as these contribute to this end.
While value reflects human preference – not only wants of the body but also what one later
writer called “wants of the mind, most of them proceeding from imagination” – price reflects
not only demand but also supply, and that is influenced by scarcity. As Barbon wrote in 1690,
“things may have great virtues, but be of small value or no price if they are plentiful”.
These quotations from the 16th and 17th centuries demonstrate an awareness of three key
principles. First, demand is separate from supply. Demand indicates what things are worth to
people; supply indicates what things cost. Second, market price reflects the interaction of both
demand and supply and, in principle, is separate from each of them. Third, the value that people
place on an item (their demand for the item) inevitably reflects their subjective preferences5 .
Returning to Adam Smith, given his distinction between value in use and price (value in
exchange), which is fundamentally the more useful measure of value? Here I part company
from Smith because, following Hume and Locke before him, he associated the true value of
an item largely with its cost of production. This English School held that, while the market
price of an item at any particular point in time is determined by demand and supply, in the long
run this will tend towards what Hume called a fundamental price, and Smith a natural price,

2
Moreover, as I shall argue, Smith’s analysis of this example is incomplete, and the conclusion he drew
from it is largely incorrect.
3
Plato Euthydemus, as cited in Bowley (1973). The discussion that follows draws on Bowley, Schumpeter
(1954), Gordon (1975), Pribram (1983) and Niehans (1990).
4
Aristotle, Politics, Book I, 9.
5
This applies to water, too, even though it is obviously an essential want of the body. Without wishing
to demean the importance of water, I will present some evidence below that, compared to other items
they could buy, people sometimes place a lower value on improving their access to water than what the
public health professionals would recommend.
64 W.M. Hanemann

which is determined by the underlying cost of production. Implicitly, they were assuming a
horizontal long-run supply curve, so that consumer demand has no influence on price in the
long run. This is now seen as a special case, and the modern economic concept of value focuses
essentially on value in use.
The modern concept was first formulated by Dupuit (1844) and Marshall (1879, 1890).
Dupuit stated that the “maximum sacrifice expressed in money which each consumer would
be willing to make in order to acquire an object” provides “the measure of the object’s utility”.
Marshall used a very similar formulation; he defined the “economic measure” of a satisfaction
as “that which a person would be just willing to pay for any satisfaction rather than go without
it”. These definitions highlight the distinction between demand and supply: the measure of
value is what the item is worth to the individual, not what it costs. Thus, an item can be cheap
to produce, in the sense that its total cost is low, but highly valuable to the owner, in that its
total value to him is large, or conversely.
Generalizing from this, the modern economic concept of value is defined in terms of a
trade-off. When an economist states that, for some individual, X has a value of 50 in terms
of Y , this means no more, and no less, than that the individual would be willing to exchange
X for 50 units of Y. Y is said to be the numeraire in terms of which value is measured. This
numeraire can be money but it need not be; it could, for example, be some specific commodity.
The trade-off is in no way limited to market goods; it can be between any two items that the
individual values, regardless of whether these have a market price.
Before any further discussion of the relation between value and price, it is necessary to
introduce a distinction which was lacking in Smith’s analysis but was understood by Dupuit
and Marshall, namely the distinction between marginal, on the one hand, and average or total,
on the other. Thus, marginal value needs to be distinguished from average or total value, and
marginal cost from average or total cost. The marginal quantity measures the change in total
value, or total cost, associated with a unit change in quantity, while the average measures total
value, or total cost, averaged over the total quantity. Admittedly, there is one case where they
are the same: if the marginal value (or marginal cost) is constant as quantity changes, then
marginal cost (or value) coincides with average cost (or value). But, in general, marginal value
and marginal cost are not likely to be constant. In particular, the general presumption is that
marginal value (and marginal utility or marginal benefit) decline with quantity.
The notion of declining marginal utility was the cornerstone of Dupuit’s analysis. Dupuit
recognized that if the consumer is free to vary the quantity of an item purchased, she will
choose this quantity so as to equate her marginal value (utility) for the item to its price. In that
case, the market price provides an accurate measure of the marginal value associated with the
last unit of consumption. But, Dupuit stressed, the total payment does not accurately reflect
the total value of all units consumed. This is because of diminishing marginal utility: if the
marginal value of the last unit just equals the market price, it follows that the marginal utility
associated with the infra-marginal units will be higher than this market price. In effect, the
consumer earns a profit on the infra-marginal units because they are worth more to her than
the price she pays, which in fact is why she consumes a larger quantity thereby rendering these
units infra-marginal. Marshall, who independently formulated a similar argument thirty years
later, called this profit the consumers’ rent in 1879, and the consumer’s surplus in 1890.
In summary, if there is a market price for the item in question and if the consumer is free to
vary the quantity of this item that she purchases, its marginal value to her is reflected in, and
can be measured by, the market price; otherwise, not. Even when price reflects marginal value,
The economic conception of water 65

total expenditure does not reflect total value; instead, total expenditure understates total value
because of the presumption of diminishing marginal utility. The distinction between marginal
and total is the key to the full resolution of the diamond and water paradox: water may have a
smaller value than diamonds at the margin, but it undoubtedly has a larger total value.
Although Dupuit and Marshall correctly enunciated the economic concept of value in its
modern formulation, it actually dropped out of favor with economists around the turn of the
last century. Marshall himself came to be troubled that his use of the demand curve to measure
consumer’s surplus was inexact and relied on the assumption of a constant marginal utility of
income. And, as the ordinal utility revolution took hold in economics, Marshall’s analysis based
on cardinal utility appeared hopelessly out-dated and irrelevant. It took until the 1970s before
these issues were fully resolved and the Dupuit-Marshall concept was recognized as being
both fully consistent with modern, ordinal utility theory and susceptible of rigorous empirical
measurement. This came about as a result of several important conceptual advances.
First, Hicks rehabilitated the Marshallian concept of consumer’s surplus in a series of papers
starting with Hicks (1939), which demonstrated that this concept is in fact consistent with
ordinal utility theory and that it could be measured exactly if one were given an indifference
map. However, this was a pyrrhic victory because the general view was that, while it is a useful
theoretical construct, the indifference map is not itself directly observable. Hence, Marshall’s
measure as re-interpreted by Hicks was not measurable in practice. This view finally changed
around 1970 as the result of the development of what is known as duality theory, including the
demonstration by Hurwicz & Uzawa (1971) of a theoretically rigorous yet practical numerical
procedure for identifying the specific utility function underlying any given system of demand
equations that satisfies the formal requirements of modern ordinal utility theory. This now
made it possible to start with an econometric estimate of a suitably specified demand equation
for a marketed commodity, or a system of demand equations for a set of commodities, and
derive a theoretically consistent and rigorous estimate of the Dupuit-Marshall measure of the
economic value of these commodities.
Second, building on Hicks (1939), Henderson (1941) discovered an alternative way of
characterizing the trade-off that underlies the economic concept of value. When one says that
a person is willing to exchange X for 50 units of Y , this could mean either: 1) the person would
be willing to give up (pay) 50 units of Y to obtain X ; or 2) the person would accept 50 units of
Y to forego X . The first uses maximum willingness to pay (WTP) as the measure of value, and
is the measure mentioned by Dupuit and Marshall and analyzed by Hicks (1939, 1941). The
second is the new measure that was suggested by Henderson; it uses minimum willingness to
accept (WTA) as the measure of value. Together, these exhaust the logically possible ways of
expressing a trade-off. Hicks (1942, 1943, 1946) analyzed the relationship between them in
the case of a price change and showed that they differ by an income effect6 .
The third development was the extension of the economic concept of value to a broader
class of items than market commodities. In fact, nothing limits X in the definition of economic
value given above to being a market good; it could actually be anything from which people
derive satisfaction. This suggests that the same definition of economic value can be applied to

6
An important paper by Willig (1976) showed how one could use Hurwicz and Uzawa’s result to develop
a tight numerical bound on the possible difference between WTP and WTA in the case of a price change
for a marketed commodity. The Hurwicz-Uzawa result implies that both WTP and WTA can be derived
from an econometric estimate of suitably specified demand equations.
66 W.M. Hanemann

non-market items. For example, one could say that a person values some aspect of his health
at 50 units of Y if he would be willing to exchange 50 units of Y to preserve that aspect of his
health; that he values a beautiful sunset at 50 units of Y if he would be willing to exchange
50 units of Y to experience it; or that he values an endangered species of animal at 50 units
of Y if he would be willing to exchange 50 units of Y to ensure its preservation. In each
case, it should be evident that there are two possible ways to formulate the exchange: a WTP
formulation and a WTA formulation. This was demonstrated formally by Maler (1971, 1974)
who showed that, when Y is money, the Hicksian analysis and its modern formulation in terms
of duality theory carry over from the valuation of market goods to non-market items7 . Maler’s
analysis thus provides a formal justification for the field of non-market valuation, including
the monetary evaluation of the natural environment.

2.2 Non-market valuation and water


Economic valuation deals with the valuation in monetary terms of items that people might
care for. Non-market valuation applies the same notion to items that are not sold in a market.
It is important to emphasize that the Dupuit-Marshall concept of economic value carries over
to such items. This is because, even for something that is not sold in a market, it is still
meaningful to conceptualize the economic measure of the satisfaction from the item as the
monetary amount which the person would be just willing to exchange for the item if it were
possible to make such an exchange. In effect, this generates a monetary measure of the change
in the person’s welfare by using the change in the person’s monetary income that she would
consider equivalent to the item in question in terms the overall impact on her satisfaction8,9,10 .
The history of non-market valuation in the USA is closely intertwined with water projects,
since these were an important motivation for the development of cost-benefit analysis. The
idea of cost-benefit analysis originated in the USA, in Hammond’s (1960) phrase, as “an
administrative device owing nothing to economic theory” in the context of managing the
activities of the US Army Corps of Engineers around the beginning of the last century. The
1902 River and Harbor Act had created a Board of Engineers to review navigation projects;

7
However, there is an important difference. With valuation of non-market items, the difference between
the WTP and WTA measures involves not only an income effect but also a substitution effect (Hanemann,
1991).
8
The equivalence can be conceptualized in two possible ways – the maximum amount that the person
would be willing to pay to gain the item, or her minimum WTA to forego it.
9
It should be noted that this definition provides a unified approach to welfare measurement for both
firms and households. In the case of households, whose objective function is defined in terms of utility,
the monetary measure is the change in income that is considered equivalent to the change in utility. In
the case of firms, whose objective function is defined in terms of profit, the monetary measure is the
change in profits itself.
10
Although the modern economic concept of value is defined in terms of a trade-off, it is possible that
some people find themselves unable to make a trade-off because for them the two items being compared
are incommensurable. A type of preference that gives rise to such trade-off aversion is where there is a
lexical ordering over commodities: certain goods in any quantity or quality always take precedence over
all quantities or qualities of other goods, so that no amount of increase in the latter can ever compensate
for any reduction in the former. This is known as lexicographic preferences. A modified version of
lexicographic preferences is where the lexical ordering applies only below a threshold level of the good
(Lockwood, 1996); among other things, this generates a situation where the individual might have an
infinite WTA for a reduction in an item.
The economic conception of water 67

in conducting a review, the Board was required to consider the commercial benefits from
such projects in relation to their costs. The River and Harbor Act of 1920 further required the
separate reporting of special, or local, benefits as opposed to general, or national, benefits
for the purpose of ensuring proper local cost-sharing. In 1934, the National Resources Board
appointed a Water Resources Committee to consider “the development of an equitable system
of distributing the cost of water resource projects, which should include not only private but
also social accounting”. Finally, the Flood Control Act of 1936 permitted the Army Corps of
Engineers to involve itself in flood control provided that, in a famous phrase, “the benefits to
whosoever they may accrue are in excess of the estimated costs”. In 1946, a Subcommittee
on Benefits and Costs of the federal Inter-Agency River Basin Committee was appointed to
investigate the practices of the various federal agencies that were engaged in the evaluation
of federal water resource projects and to formulate some “mutually acceptable principles
and procedures”. This led ultimately to the publication in 1950 of what became known as
the Green Book which attempted to codify the principles of cost-benefit analysis for use by
federal agencies. The following decade saw the publication of many academic journal articles
and six major academic books dealing with the economic analysis of water projects.
The 1950s were when the field of non-market valuation began to come into existence.
The approach that emerged first is what became known as the travel cost method or, more
generally, the revealed preference method. It arose initially out of an effort by the National
Park Service (NPS) to measure the economic value associated with the national parks. At the
time there were no entrance fees at national parks, so the NPS could not use park revenues
as a measure of their value. The issue was assigned to a staff economist who wrote to ten
distinguished economic experts for their advice. All but one replied that it was impossible
to measure recreational values in monetary terms, but the tenth, Harold Hotelling disagreed.
He saw that, even though there was no entrance fee for a national park, it still cost visitors
something to use the parks because of expenses for travel, lodging and equipment. These
expenditures were not captured by the NPS but, they still set a price on the park. Moreover,
this price would vary among people coming from different points of origin. By measuring the
price and graphing it against visitation rates one could construct a demand schedule for visits
to the site, and then determine consumer’s surplus in the usual manner as the area under this
demand curve.
The NPS report followed the majority view and asserted that it was not possible to set
a monetary value on outdoor recreation. However, in 1956 the State of California hired an
economic consulting company to estimate recreational benefits associated with the planned
State Water Project. This company learned of Hotelling’s idea and decided to apply it. A survey
of visitors was conducted at several lakes in the Sierras and data were collected on how far
they had traveled and how much they had spent. Using these data, a rough demand curve was
traced out, and an estimate of consumer’s surplus was constructed. This analysis appeared in
Trice & Wood (1958), the first published application of the travel cost method. At the same
time, Marion Clawson (1959) at Resources for the Future had begun collecting data on visits
toYosemite and other major national parks in order to apply Hotelling’s method to them, which
was the second published application. By 1964, there were at least five more applications in
various parts of the USA, and the travel cost method was an established procedure.
The insight behind the travel cost method, and revealed preference generally, is that, while
people cannot buy non-market goods such as clean water or an unspoiled environment directly,
there sometimes exist market goods that serve as a partial surrogate for the non-market good
68 W.M. Hanemann

because the enjoyment of these goods is enhanced by, or depends on, the non-market good.
In that case, the demand for the market goods is used as a surrogate for the demand for the
non-market good.
The limitation of this approach is that there may not exist a market good that can serve as
surrogate for the non-market good of interest. Moreover, even if such a good exists, it may not
capture all of people’s preferences for the complementary non-market good. The conceptual
identification of what might be omitted by the revealed preference approach came about as a
result of papers by Weisbrod (1964) and Krutilla (1967). Both authors started from the premise
that some of people’s motives for valuing the natural environment may differ from those
for valuing a market good. People may value the natural environment out of considerations
unrelated to their own immediate and direct use of it. Weisbrod focused on uncertainty and
what became known as option value: some people who do not now visit a national park,
say, may still be willing to pay money to protect it from destruction or irreversible damage
because they want to preserve their option of visiting it in the future. Krutilla focused on
what became known as bequest value and existence value11 . With bequest value, the notion
is that some people would be willing to pay because they want to preserve the park for future
generations. With non-use value, the notion is that some people would be willing to pay even
if they knew that neither they nor their children would ever visit it; in Krutilla’s example,
people may “obtain satisfaction from mere knowledge that part of the wilderness in North
America remains”. These are legitimate sources of value, Krutilla and Weisbrod felt, but they
would not be respected by private managers of the environmental resource. Nor would they
be adequately measured by a conventional revealed preference analysis such as the travel cost
method. Consequently, some other method of measurement is needed.
The alternative approach, suggested by Ciriacy-Wantrup (1947), is to interview people and
elicit their monetary value; this became known in economics as the contingent valuation (CV)
method12 . Ciriacy-Wantrup was discussing soil conservation and he noted that several of the
benefits were non-market goods, such as reduced siltation of rivers or reduced impairment of
scenic resources. He characterized the problem as being how to obtain a demand curve for such
goods, and suggested the following solution: “[Individuals] may be asked how much money
they are willing to pay for successive additional quantities of a collective extra-market good.
The choices offered relate to quantities consumed by all members of a social group . . . If every
individual of the whole social group is interrogated, all individual values (not quantities) are
aggregated”. The results correspond to a market-demand schedule. While noting the possible
objection that “expectations of the incidence of costs in the form of taxes will bias the responses
to interrogation”, he felt that “through proper education and proper design of questionnaires
or interviews it would seem possible to keep this potential bias small”.
Having identified a solution conceptually, Ciriacy-Wantrup never pursued it further. The
first significant application was by Davis (1963) which dealt with the economic value of
outdoor recreation in the Maine woods; to measure this Davis interviewed a sample of hunters

11
The latter is now also called non-use value and passive use value.
12
The same idea was earlier suggested by Bowen (1943) who conceived of surveys as a surrogate for
using voting to determine the public’s demand for what he called social goods. Recently, the term stated
preference has been used to cover CV and related approaches. They are also known as direct valuation
whereas revealed preference approaches are referred to indirect valuation because they do not measure
preferences directly but instead infer them from externally observed behavior.
The economic conception of water 69

and recreationists and asked how much more they would be willing to pay to visit the area13 .
The next application was by Ridker (1967); to measure the damages from air pollution, Ridker
included some questions in a survey about people’s WTP to avoid soiling from air pollution.
In 1969, a steady stream of CV studies began to appear in the economics literature. Official
recognition was given to CV in 1979, when the US Water Resources Council included it along
with travel cost as recommended methods of non-market valuation.
The first application of non-market valuation in the USA, the 1957 valuation of recreation
benefits from the California State Water Project, was a harbinger of things to come: since
then many non-market valuation studies have been conducted in the USA in connection with
water resources management issues, and the environmental consequences of water projects
have come to play a significant role in the design and approval of water projects. These
trends emerged slowly in the 1970s and 1980s, driven by developments in the implementation
of the 1969 National Environmental Policy Act (NEPA)14 . NEPA required federal agencies
to prepare a detailed statement of environmental impacts for proposed major actions which
significantly affect the quality of the human environment, including the identification of the
environmental impacts of the proposed action, alternatives to the proposed action, and any
adverse environmental impacts which cannot be avoided should the proposal be implemented.
In consequence, since the mid-1980s it has not been acceptable in the USA to perform an
economic assessment of a major water project without including some non-market valuation
of the project’s environmental impacts. For example, non-market valuations of environmental
impacts were included in the Department of Interior’s re-assessment of the operation of Glen
Canyon Dam on the Colorado River in 1984–1992, and in the Bureau of Reclamation’s assess-
ment of the Central Valley Project Improvement Act in 1993–1996. In California, they were
included in the State Water Resources Control Board’s review of the diversions of water from
the San Francisco Bay/Delta to the Central Valley and Southern California, conducted in 1987–
1994, and in the Board’s 1993 Mono Lake Decision requiring Los Angeles to reduce its diver-
sion of water from streams feeding Mono Lake on the eastern side of the Sierra Nevada15 . In
the case of Mono Lake, the Board decided that it was in the public interest to reduce Los Ange-
les’ diversion from Mono Lake by about two thirds, despite the resulting loss of hydropower
and water supply (which amounted to over 8% of Los Angeles’ total water supply) primarily
in order to protect habitat for birds and other wildlife; non-use values associated with habitat
protection constituted the main component of environmental benefits (Wegge et al., 1996).
It should be emphasized that the use of non-market valuation applies to positive as well
as negative environmental impacts of water projects. The experience in the USA has been
that these can generate significant economic benefits associated with water-based recreation,
eco-tourism, and the non-use value of ecosystem protection. These environmental benefits
sometimes greatly outweigh the benefits from agricultural or even urban water use. In short,
in the USA we have now moved from the traditional situation where there was essentially a
single objective for large water projects, namely the provision of water for off-stream uses,

13
Probably the first CV study was actually conducted in 1958 for the NPS, which hired a market research
company to survey residents of the Delaware River basin about their WTP entrance fees for national
parks.
14
Subsequently, a number of states passed laws imposing similar reporting requirements on agencies of
the state government; for example, the California Environmental Quality Act was enacted in 1970.
15
I served as the Board’s economic staff for its investigation of both these issues.
70 W.M. Hanemann

to a situation where any new water project must have environmental restoration as an explicit
objective along with the provision of any off-stream uses.

3 IS WATER DIFFERENT?

Now that the economic concept of value has been explained, the question arises whether
it is appropriate to apply this concept to water. Is water an economic commodity, and can
it be analyzed using the conceptual framework of economics in the same way as any other
commodity?
The answer is contested ground between economists and their critics. One of the four Dublin
Principles, adopted at the 1992 International Conference on Water and the Environment in
Dublin, holds that “water has an economic value in all its competing uses and should be recog-
nized as an economic good”. Similarly, Baumann & Boland (1998) write: “water is no different
from any other economic good. It is no more a necessity than food, clothing, or housing, all of
which obey the normal laws of economics”. Per contra, Barlow & Clarke (2002) proclaim it
as a “universal and indivisible” truth that “the Earth’s freshwater belongs to the Earth and all
species, and therefore must not be treated as a private commodity to be bought, sold, and traded
for profit … the global freshwater supply is a shared legacy, a public trust, and a fundamental
human right, and therefore, a collective responsibility”. Vandana Shiva (2002) writes in a
similar vein about a clash between two cultures: “a culture that sees water as sacred and treats
its provision as a duty for the preservation of life and another that sees water as a commodity,
and its ownership and trade as fundamental corporate rights. The culture of commodification
is at war with diverse cultures of sharing, or receiving, and giving water as a free gift”.
My own view lies somewhere between these two positions. Baumann & Boland are undoubt-
edly correct when they point out that food, clothing and shelter, like water, are necessities of
life, and they are typically provided through the market without any complaint. Why, they
ask, should water be different? I believe there are two reasons why this is so. First, water is
clearly viewed by many people as being different. The fact that water, unlike other household
commodities, arouses such passion speaks for itself: for better or worse, water is perceived as
having a special significance that most other commodities do not possess16 . This itself has eco-
nomic consequences. Second, I believe that water has some other economic features that make
it distinctive. These features make water different from, say, bread or land, as an economic
commodity yet they are often overlooked by economists. They matter greatly because they
affect the demand for water, its value, and the social and institutional arrangements by which
it is supplied. To explain them, I need to introduce several more items from the economists’
conceptual toolkit.

3.1 Water as a private good, water as a public good


Since Samuelson (1954), economists have drawn a distinction between conventional market
goods – also known as private goods – and what are known as public goods, “which all enjoy

16
This is true in rich as well as poor countries – in the USA, for example, it is notoriously difficult
for publicly owned urban water utilities to obtain political approval for even trivial rate increases while
other household utilities such as cable television raise their rates with impunity; Glennon (2004) makes
a similar observation.
The economic conception of water 71

in common”. The two key properties of a public good are non-rivalry in consumption and
non-excludability. With conventional goods, one person’s consumption necessarily competes
with that of another, in that more consumption by one person renders a smaller quantity of
that good available for consumption by anybody else. With public goods, by contrast, more
consumption by one person in no way reduces the amount available for others. Conventional
consumption goods are excludable in that, if this is so desired, it is physically possible to
exclude any person from consuming the commodity. With public goods, by contrast, if the
good is available for consumption by anybody, it is available for consumption by all. Examples
of a public good suggested by Samuelson were “an outdoor circus or national defense which is
provided for each person to enjoy or not, according to his tastes”. The abatement of pollution in
a lake is another example of a public good, as are other types of environmental improvement:
my enjoyment of the clean water in the lake in no way reduces the amount of clean water
available for your enjoyment (non-rivalry) and, if the water in the lake is clean for me to enjoy,
it is clean for everyone’s enjoyment (non-excludability)17 .
In this framework, water is both a private good and a public good. When water is being
used in the home, in a factory or on a farm, it is a private good. When water is left in situ,
whether for navigation, for people to enjoy for the view or for recreation, or as aquatic habitat,
it is functioning as a public good18 . Moreover, while the water in a reservoir is a private good,
the storage capacity of the reservoir per se may be a public good. By contrast, most of the
other commodities associated with food, clothing or shelter are purely private goods and have
no public goods aspect; this is one of the respects in which water is different than these other
commodities in economic terms.
Samuelson identified two important consequences of the public good properties. First,
while public goods are likely to be supplied collectively, for example through a voting process,
rather than through a decentralized market, it is likely that they will be undersupplied because
people have a selfish incentive to free ride on the collective decision process by understating
their true interest in the public good. Second, the valuation of public goods is fundamentally
different than that for private goods because a public good can be enjoyed simultaneously by
many while a private good can be consumed by only one party at a time. Thus, the value placed
on a given unit of a private good is that of a single user – in an efficient market, this will be the
user with the highest and best use for the item. By contrast, the value placed on a public good
is that of many people, namely all those who care for the item19 . This is why the non-market

17
In addition to private and public goods, there is an intermediate case where there is rivalry in con-
sumption but not excludability. These are known as common pool resources. Examples include fisheries,
forests, grazing grounds, and oil fields. The other intermediate case, sometimes called club goods or
quasi-private goods, is where there is non-rivalry combined with the possibility of exclusion. Examples
include television frequencies, public libraries, and bridges, for each of which it is possible to exclude
access. Furthermore, there may be non-rivalry at low levels of aggregate consumption of a club good, but
rivalry at a high level of consumption once the item becomes congested – this can happen, for example,
with parks and bridges.
18
To the extent that water-based outdoor recreation is excludable, this would be a quasi-private good. To
the extent that groundwater or water flowing through the distribution system of an irrigation district is
non-excludable, these are common pool resources.
19
This follows from Samuelson’s demonstration that the aggregate demand curve for a public good is
constructed in a radically different manner than the aggregate demand curve for a private good. With
a private good, the aggregate demand curve is the horizontal sum of every individual’s demand curve
for the good; with a public good, the aggregate demand is the vertical sum; this observation had in
72 W.M. Hanemann

benefits of environmental preservation can sometimes outweigh the use benefits associated
with the diversion of water for off-stream agricultural or urban use.
The public good nature of water in situ, historically associated with navigation, has had a
decisive influence on the legal status of water. In Roman Law and, subsequently, in English
and American common law, and to an extent in Civil Law systems, flowing waters are treated
as common to everyone (res communis omnium), and are not capable of being owned. These
waters can only be the object of rights of use (usufructuary rights), but not of rights of
ownership20 . Thus, even though water and law are often complementary inputs, there is a
crucial distinction in that land can be owned, while water cannot.

3.2 The mobility of water21


A distinctive physical feature of water is its mobility. Water tends to move around. It flows, it
seeps, it evaporates. When water is applied to plants in the field (or to an urban landscape), a
substantial portion either seeps into the ground or runs off the ground as tailwater. In addition,
in residential indoor uses and most industrial uses there is usually an outflow of wastewater
after the use is completed. The consequence is that there can be several sequential uses of the
same molecule of water since water is rarely consumed fully by a given user and what is left
is physically available, in principle, for use by others22 .
The mobility of water and the opportunity for sequential use and re-use make water relatively
distinctive as a commodity – especially compared to land, for which such multiple, sequential
uses are impossible (except in nomadic societies). These properties of water have important
economic, legal and social implications. Keeping track of water flows is costly and sometimes
difficult. Consequently, it is often hard or impractical to enforce excludability or to establish
property rights to return flows. In this respect, water is very different as an asset than land,
which is relatively easy to divide and fence. The common solution is to resort to some form
of collective right of access; in effect, this internalizes the externality associated with the
mobility of return flows. A classic example of this is the riparian water right in English and
American common law. This permits any landholder whose property is adjacent to a stream
or body of water to divert a reasonable amount of water, provided this does not cause harm to
other riparian landholders or interfere with their co-equal right to divert a reasonable amount
of water. The riparian right to the use of water is not a right to a fixed quantity, and it is a
co-relative right shared with all other riparians along the same stream23 .

fact already been made by both Bowen (1943) and Ciriacy-Wantrup (1947). In terms of the distinction
between use and non-use values, if there is a non-use value for an item this is a public good.
20
By contrast, groundwater beneath private land and springs or rainwater found on private land is
typically treated by the law as being privately owned by the landowner(s) on whose land the spring
occurs or under whose land the groundwater lies.
21
The analysis in this and the following two sections was influenced by reading Young & Haveman
(1985).
22
In the process, however, there can often be some reduction in the quality of the water relative to that in
the first use. Because of the solvent properties of water, the return flows are apt to dissolve and absorb
chemicals in the media through which they pass.
23
In American law, there is a further requirement that riparians put the water to a reasonable use. With
surface water, the major alternative in American law to the riparian right is the appropriative right, which
was developed in the arid West. This permits the diversion of water, regardless of whether the diverter
owns the riparian land, in a fixed quantity, subject to the principle of first in time is first in right. The
The economic conception of water 73

3.3 The variability of water


In addition to the mobility of water in streams, another crucial feature is the variability of
supply in terms of space, time, and often quality. Spatially, water is distributed very unevenly
across much of the globe; just six countries – Brazil, Russia, Canada, Indonesia, China and
Colombia – account for half of the world’s total renewable supply of freshwater (Postel &
Vickers, 2004). Even within countries and regions, there is unequal spatial distribution. In
California, for example, two thirds of the state’s population live in Southern California, but this
region receives less than 10% of the state’s total precipitation. For any given region, there is
substantial variation in precipitation both within the year and between years. In California, for
example, approximately 80% of the annual precipitation falls between October and March,
while three quarters of the water use occurs between April and September24 . Beyond this,
cycles of wet and dry years occur in California as a function of wider climatic phenomena
such as the interannual El Niño-Southern Oscillation and the Pacific Decadal Oscillation.
While the annual runoff in California has averaged about 87,500 Mm3 over the past 90 years,
it has been as low as 18,500 Mm3 (in 1977) and as high as 166,500 Mm3 (in 1983).
Because of this variability, the major challenge for most large water systems is the spatial
and temporal matching of supply with demand. Storage is typically the key to controlling
the temporal variability in supply, while inter-basin transfers are used to overcome the spa-
tial mismatch between supply and demand. But, the variability of supply has affected not
just the engineering of water resource systems but also the legal, and institutional arrange-
ments for the use of water. The variability of supply is yet one more point of divergence
between water and land, and it explains why the property rights regimes are different: it
would surely be difficult to apply the ownership rights in land to so variable a resource as
water.
Besides the variability in supply, the demand for water may be intermittent, especially in
agricultural uses of water where crops need to be irrigated only at periodic intervals rather
than every hour of every day. Until the advent of affordable storage, which has mainly been
a phenomenon of the 20th century and large-scale diversion of water, the intermittent nature
of traditional agricultural demand was an important factor promoting the sharing of access.
If there is water in the stream and one member of the group is not currently diverting water

theory is that, if the streamflow is inadequate to meet all the diversion requirements, those with a more
recent (junior) date of initial diversion cede to those with an older (more senior) date. On the ground
in at least some Western states, the practice seems to be rather different. The precedence of seniority
is not self-enforcing without resort to litigation, which is slow and costly. Much of the time, therefore,
what actually happens with appropriative rights may be closer to a version of the riparian system. In
California, there is not a functioning system to record the actual diversions of water, nor to check these
against the quantity associated with the water right. Consequently, much of the surface water use by
agricultural occurs outside the formal structure of California appropriative water rights law. This can
become an impediment to long-run water transfers as the inadequate documentation casts a shadow of
doubt on a seller’s specific property right.
24
The seasonal variability of precipitation in California is exacerbated by the fact that it is an arid region
with a Mediterranean climate. But precipitation is distributed unevenly throughout the year in almost all
parts of the world, albeit not as severely as in California. In Europe, the major part (46%) of the runoff
occurs during April to July, and similarly in South America; in Asia, 54% occurs in June to September;
in Africa, 44% occurs in September to December; in Australia and Oceania, 40% occurs in January and
April (UNESCO, 2000).
74 W.M. Hanemann

from the stream, other members of the group were allowed to divert the water rather than let
it flow to waste in the ocean25 . This is another key difference with land: while the demand for
water is intermittent, the demand for land to grow crops or to locate a building is continuous,
and there can be no such sharing of the same resource among multiple users. The intermittent
nature of the agricultural demand for water is conducive to the collective sharing of a right of
access as opposed to individual ownership of a property right.

3.4 The cost of water


Compared to other commodities, and other utility services, the cost of water has several
distinctive features which complicate its supply.
Water is bulky, and expensive to transport relative to its value per unit of weight. Con-
sequently, the transportation infrastructure for water is far less extensive than that for more
valuable liquids such as petroleum. Also, compared to electricity, water is relatively expen-
sive to transport, but relatively cheap to store. Therefore, the strategy for averting shortage
takes a different form with water than electricity. If there is a sudden shortfall in supply, with
electricity this can be made up almost instantaneously by importing power over the grid from
a source that could be 1500 km away or more. With water, there is no comparably intercon-
nected transportation grid and, even if there were, it takes longer to move a comparably large
quantity of water. Thus, to deal with unexpected outages, one has to either resort to rationing
or stockpile sufficient stored water prior to the period of peak use.
Another distinctive economic feature is that water supply is exceptionally capital-intensive
compared not only to manufacturing industry generally but also to other public utilities. In the
USA, for example, the ratio of capital investment to revenues in the water industry is double
that in natural gas, and 70% higher than in electricity or telecommunications. Moreover,
the capital assets used in water supply cannot be moved to another location and are generally
unusable for any other purpose; they represent an extreme type of fixed, non-malleable capital.
Furthermore the physical capital in the water industry is very long-lived. The infrastructure
associated with surface water storage and conveyance and the pipe network in the streets can
have an economic life of 50–100 or more years, far longer than that of capital employed in
most manufacturing industry or in other public utility sectors26 .
In addition, there are significant economies of scale in many components of water supply
and sanitation. These are especially pronounced for surface water storage: given a specific

25
Storage changes this, because streamflow can be stored when it is not currently being used. The
discussion here focuses on agricultural rather than urban use of water. Urban use is different because it
is more continuous in nature, and when there is a piped water supply this is typically pressurized, unlike
with agriculture which relies mainly on gravity flow. Gravity distribution fosters sharing of intermittent
access, while pressurized distribution fosters simultaneous individual access.
26
The Roman aqueduct that still stands in Segovia, Spain, is ample testimony to the physical longevity of
certain types of conveyance structure. The effective life of a dam is governed by the rate of siltation but can
be well over a century. For piping, the American Waterworks Association recommends a replacement
cycle of 67 years. A conventional drinking water treatment plant may have a useful life of 40 years,
although high technology processes such as reverse osmosis facilities have a shorter life. Compared
to surface water, the capital infrastructure associated with groundwater typically has a shorter life; a
groundwater pump might typically have a life of about 25 years in the case of an electric pump, or about
15 years for a diesel pump.
The economic conception of water 75

dam site, within some range, by increasing the capacity of the dam one can significantly
reduce the unit cost of stored water. With a groundwater source, by contrast, the economies of
scale in production are much less pronounced. There are also important economies of scale
in the treatment and conveyance of drinking water and wastewater27 .
The capital intensity, longevity, and economies of scale mean that water supply and sani-
tation costs are heavily dominated by fixed costs. In a simple surface-water supply system
with minimal treatment of drinking water, minimal treatment of sewage prior to discharge, and
a heavy reliance on gravity flow, the short-run marginal cost of water supply and sanitation
may be almost zero except for small costs associated with pumping to move water through
the system28 . Even in a modern system with full treatment of drinking water and sewage
discharges, the short-run marginal costs are extremely low. There is thus an unusually large
difference between short- and long-run marginal cost in water supply29 .
The capital intensity and economies of scale associated with surface water supply have
profound economic and social implications. For one thing, because these are classic pre-
conditions for a natural monopoly, they make it more likely that there will be a single provider
in any given area. More generally, they foster public provision of a surface water supply rather
than individual, self-provision, whether the public provision is by a collective of the users or
a monopoly seller30 . Furthermore, the construction and operation of large-scale surface water
storage and distribution systems require a high degree of co-ordination and social control. This
was the central thesis of Wittfogel’s (1957) study of ancient hydraulic societies – civilizations
that were dependent on large-scale surface water diversion and distribution systems, such
as Mesopotamia, Egypt and China. Wittfogel argued that, in such societies, the effective
provision of water required the centralization of power and an oriental despotism mode of
governance in which a state bureaucracy, headed by an absolute ruler, ruled on the basis of
its control of the hydraulic system. Wittfogel’s work was subsequently criticized by other

27
With the conveyance of drinking water from the point of production to the point of use, and of treated
wastewater to the point of discharge, there are economies of scale with respect to volumetric capacity
but not length.
28
In the USA, the ratio of operating costs to total costs for efficient water firms is about 10%; by contrast,
it is 32% for gas utilities and over 57% for electric utilities (Spiller & Savedoff, 1999). In the UK water
industry, Armstrong et al. (1994) report that operating costs represent less than 20%, and fixed costs
more than 80%, of total costs.
29
When piped water supply was introduced into cities in the 19th century, water agencies chose not to
meter individual homes or small non-residential users partly because of an ethos in favor of promoting
universal service but also because water was so cheap at the margin that they felt it was not worth the cost
of metering it. By contrast, electricity and gas were metered in residential connections. Water service
was financed by charges based on the type or value of the property being served. In the USA, metering of
residential users did not become common until well into the 20th century, and there were some notable
handouts (Denver and New York City did not meter until about 15 years ago). In the UK, nearly all
residential water users were unmetered prior to privatization in 1990.
30
For urban water, the main alternatives to a public supply of surface water are water vendors or household
self-provision through pumped wells or rainwater catchments. Where there are water vendors, the unit
cost of vended water is always much higher than that of water from piped supply, typically by a factor of
10 or more. The primary reason for the cost differential is economies of scale: it is far more expensive to
deliver water in relatively small quantities through multiple trucks rather than in large quantities through
a single pipe network. However, the up-front capital investment required for vended water is far lower
than for a piped water supply, and this can make vended water a viable alternative.
76 W.M. Hanemann

scholars31 . Nevertheless, his notion still resonates; it was applied to the American West by
Worster (1986), who characterized this as a hydraulic society based on the development and
control of water infrastructure by a political elite. This characterization is contested by Kupel
(2003), who argues that the history of water projects in Arizona is one of response by civic
leaders to requests for service by urban and suburban residents, closely resembling other
aspects of the history of modern urban infrastructure. This is not necessarily a contradiction:
the commonality in urban infrastructure is capital intensity and economies of scale, with the
consequent need for public sector leadership and social co-ordination and, also, the consequent
prospect of a handsome increase in land value in the area being served32 .
Another problematic consequence of the capital intensity, longevity of capital, and
economies of scale in surface water infrastructure is the propensity to what might be called
lumpiness or, less politely, gigantism in these systems. Because of the economics, there is a
strong incentive to make a substantial expansion of capacity at a single point in time rather
than to plan for a series of incremental changes spread out over time33 . The drawback is that it
may take many years, or decades, before the demand materializes to utilize this capacity (and
the willingness to pay – WTP – to finance it). When fully utilized, the project provides water
at a low cost; but there is uncertainty whether and when it will be fully utilized, and meanwhile
it ties up scarce capital. Large surface water projects are risky, and difficult, inter-temporal
balancing acts34 .

3.5 The price of water


It is important to emphasize that the prices which most users pay for water reflect, at best,
its physical supply cost and not its scarcity value. Users pay for the capital and operating
costs of the water supply infrastructure but, in the USA and many other countries, there is
no charge for the water per se. Water is owned by the state, and the right to use it is given

31
It was pointed out that large-scale irrigation works in Mesopotamia were developed after the rise of a
centralized state, so that hydraulic society could be the result rather than the cause of state formation. The
Maya civilization, where irrigation was of marginal importance, was cited as evidence that centralized
states might not always be associated with hydraulic systems. It was also noted that there are several
modern communities in Mesopotamia where small-scale cooperative irrigation works without centralized
external control.
32
As noted above, the increase in land value made it possible to finance urban water infrastructure with
property taxes rather than through user charges. This may have been economically rational not only
because of the very low marginal cost of urban water, but also because of the public good benefits of
urban water supply associated with improved fire protection and also what, in the 19th century, were
believed to be the public health benefits of washing down streets (Anderson, 1980).
33
By contrast, systems supplied from groundwater are considerably less lumpy and more scaleable.
34
This is well illustrated by the experience of the Central Arizona Project, the most recent large water
project in the USA, which actually went bankrupt (Hanemann, 2002). The two key parameters for the
economic viability of a water project are the discount rated use to assess the present value of net benefits,
and the rate of growth in the public’s ability and, more importantly, WTP for the water. Public agencies
can generally borrow at a lower interest cost than private firms, and also are more apt to take a long-term
perspective in evaluating investment. The public sector’s ability and willingness to apply a low discount
rate is a major reason for its predominant role in the provision of water supply infrastructure. For all the
rhetoric on privatization of water, the private sector seems more interested in taking over the operation
of existing infrastructure rather than financing new infrastructure, except for water treatment facilities
which, as noted above, have a shorter life than other water supply infrastructure.
The economic conception of water 77

away for free. Water is thus treated differently than oil, coal, or other minerals for which the
USA government requires payment of a royalty to extract the resource. While some European
countries, including England, France, Germany and Holland, do levy an abstraction charge
for water, these charges tend to be in the nature of administrative fees and are not generally
based on an assessment of the economic value of the water being withdrawn. Thus, in places
where water is cheap, this is almost always because the infrastructure is inexpensive, or the
water is being subsidized, rather than because the water per se is especially abundant.
In the USA, it has long been noted that the prices charged to farmers are far lower than
those charged to urban residents, often by a factor of 20 or more. It is often assumed that this
is because the irrigation water has been subsidized by the federal government, but this is not
in fact the main reason.
It certainly is true that the federal government has subsidized irrigation in the West by
waiving interest charges and other means. Between 1902 and 1994, the federal government
spent US$ 21,800 million to construct 133 water supply projects in the West. Although most of
the water from these projects is used for irrigation, the cost allocated for repayment by irrigation
users was set at US$ 7100 million (33%). Of this, US$ 3700 million was subsequently waived.
Of the remaining US$ 3400 million payable by irrigators, only US$ 950 million had actually
been repaid as of 1995 (General Accounting Office, 1996). The remaining balance will not be
paid off until well into this century, if at all. The combined effect is that recipients of irrigation
water from federal projects will have repaid, on average, about US$ 0.10 on each dollar of
construction cost. However, these projects account for only about 19% of total irrigation supply
in the West. The remainder comes from groundwater or non-federal surface supply projects,
none of which is subsidized to a significant degree35 . While the non-federal irrigation supply
is more expensive than the federal supply, it still is much cheaper than urban water supply.
The reason is the sharp difference in the real cost of agricultural versus urban water supply.
Unlike urban water, irrigation water is not treated, and it is generally not available on demand
via a pressurized distribution system. Moreover, the physical capital used for irrigation supply
is often old and long-lived, and it may have been paid off long ago.
There is an additional tendency to under price water in the USA – urban as well as agri-
cultural – because most water agencies set price to cover the historic (past) cost of the system
rather than the future replacement cost36 . There is typically a large gap between these two costs
because of the extreme lumpiness and longevity of surface water supply infrastructure. The
capital intensity of the infrastructure exacerbates the problem because, after a major surface
water project is completed, since supply capacity so far exceeds current demand, there is a
strong economic incentive to set price to cover just the short-run marginal cost (essentially, the
operating cost), which is typically minuscule37 . As demand eventually grows and the capacity
becomes more fully utilized, it is economically optimal to switch to pricing based on long-
run (i.e. replacement) marginal cost, but by then water agencies are often politically locked

35
If there is a subsidy, it is likely to be mainly for the electricity used in pumping water.
36
This is due partly to the conservatism of the conventional engineering emphasis on cost recovery,
and partly to the fact that, since most water supply agencies in the USA are publicly owned rather than
investor-owned, there is a strong ethos to avoid making a profit on the sale of water.
37
Erie & Joassart-Marcelli (2000) argue that this type of water pricing encouraged urban growth, and
urban sprawl, in Southern California, but they fail to recognize the economic logic that drives it by virtue
of the lumpiness and capital intensity of water supply infrastructure.
78 W.M. Hanemann

into a regime of low water prices focused narrowly on the recovery of the historical cost of
construction.

3.6 The essentialness of water


Water is essential for all life – human, animal, or plant. In economics, there is a concept,
also called essentialness, that formalizes this notion. The concept can be applied either to
something that is an input to production or to something that is directly enjoyed by people as a
consumption commodity. In the case of an input, if an item has the property that no production
is possible when this input is lacking, the item is said to be an essential input. In the case of
a final good, if it has the property that no amount of any other final good can compensate
for having a zero level of consumption of this commodity, then it is said to be an essential
commodity. Water obviously fits the definition of an essential final good: human life is not
possible without access to 5 or 10 L/d of water per person. Water also fits the definition of an
essential input in agriculture and in several manufacturing industries (e.g. food and beverages,
petroleum refining, lumber and wood products, paper, chemicals, and electronic equipment)
that cannot function without some input of water38 .
However, essentialness conveys no information about the productivity or value of water
beyond the vicinity of the threshold. It implies nothing about the marginal value associated with,
say, applying 76 versus 89 cm of water to irrigate cotton in the Central Valley of California.
It says nothing about the marginal value of residential water use at the levels currently expe-
rienced in Western Europe or the USA – the latter averages about 455–530 L/d per person,
more than two orders of magnitude larger than the minimum quantity that is needed for human
survival39 .
The latter statement is not meant to belittle the uses of water by households in Western
Europe or the USA. My point is that, in addition to being essential for human life, water
contributes in important ways to the enjoyment of the satisfactions of life. Consequently,
there are many other residential end uses of water besides its use for drinking. Indeed, if one
examines the history of residential water use in the USA from the early 19th century to the
present, it is striking how water consumption has grown over time through the steady accretion
of end-uses, each representing the discovery of a new way to employ water for people’s use
and enjoyment. When a piped water supply first became available in the 19th century, the
initial household uses were the same ones that had existed when family members had to fetch
water from an external source – drinking, cooking, hand washing, and limited bathing40 .
As time passed, many other uses were found – tubs for bathing, water borne sanitary waste
disposal, outdoor landscape and garden watering, automatic clothes washers, swimming pools,
automatic dish washers, car washing, garbage disposal, indoor evaporative cooling, hot-tubs,
lawn sprinklers, etc. The result has been a constantly rising trajectory of per capita household
water use.

38
These are the largest water-using industries in the USA in terms of freshwater intake.
39
Total urban water use in the USA averages about 680–830 L/d per capita, depending on the location.
40
For example, Blake (1956) notes that out of 15,000 houses with running water in Philadelphia in 1849,
only about 3500 were equipped with private baths. In 1871, by contrast, 112,457 Philadelphia houses
had running water, and 80,000 of these had bathtubs and fixtures for hot and cold water (Anderson,
1980). The use of water closets to remove human wastes did not become widespread until almost two
decades after the introduction of piped water into homes (Tarr, 1979).
The economic conception of water 79

Two conclusions can be drawn from this historical experience. First, in developed countries,
the fact that water is essential for human life is almost certainly irrelevant when assessing
the value of residential water supply because the ways in which water is used are nowhere
near the threshold level at which essentialness applies. Second, there is a possibility that
some of the developing countries may also move along a rising trajectory of residential water
consumption because the things that have made abundant water use an element of a comfortable
modern life style in developed countries could also become attractive to people in these
countries as their income rises.
For reasons that are entirely understandable, there is a tremendous emphasis in the water
literature on the need to secure at least a minimal water supply for the nearly 1100 million
people around the globe who currently live without access to an improved water source. In
developing this estimate, the United Nations and the WHO used a figure of 20 L/d per person
as the minimum human requirement for water for drinking and basic sanitation. Gleick (1999)
has argued that this is too low, and has advocated the adoption of a basic human right to 50
L/d per person as the minimum required for bathing and cooking as well as the other basic
needs. As noted above, these low levels of consumption are not very different from the initial
levels of water use in developed countries when piped water was first introduced in the 19th
century. It is important to recognize that, if the efforts to provide improved water supply and
sanitation in the developing countries are successful, the future levels of water use that might
ultimately emerge in these countries could diverge from these minimum levels, as happened
in the developed countries.
In short, while it is obviously appropriate to think in terms of human needs for water, it is
also appropriate to recognize that people also have demands for water as a commodity that
generates pleasure by utilizing it in various ways. As they become more affluent, poor people
are likely to choose to allocate more of their resources to satisfying not just their needs of the
body but also their “wants of the mind, most of them proceeding from imagination”, perhaps
including some domestic end uses of water that might seem outlandish today.
If and when this broadening of domestic water use beyond the basic minimal level occurs,
an important implication is that planners will need to adopt a behavioral approach to the anal-
ysis and projection of urban demand, as opposed to the engineering/public health approach
that dominates the literature on water and poverty today. The behavioral approach focuses
not on how much water people need but rather on how much water they are willing to
pay for.
The difference between the engineering and behavioral perspectives is well illustrated by
the experience of the World Bank over the past fifteen years. Water planners in the Bank
originally thought that water and sanitation projects in developing countries were not viable
if they required households to pay more than 3–5% of their income for the project services,
because this would be more than they would be willing or could afford to pay, rendering
the projects infeasible (World Bank, 1975). It became evident from detailed household-level
studies sponsored by the Bank in the 1980s and 1990s that, in many developing countries,
some households spend considerably more than this on access to traditional, unimproved water
and sanitation. Some households purchase water from vendors at prices which can be much
higher than the cost of piped water. Where there is piped water supply, many households incur
expenses on installing storage capacity in the home to ensure that they have water when the
pipes run dry; others undertake a wide variety of practices to treat contaminated water in
their home to make it safe to drink. Moreover, some carefully designed contingent valuation
80 W.M. Hanemann

surveys by Dale Whittington and his colleagues showed that some households have a WTP
for improved water supply that can exceed 3–5% of household income41 .
The issue is not how much a household values access to water versus no access to water
at all but, rather, how much it values a piped, public water supply relative to the existing
alternatives. In this context, it is interesting to review what is known about how the adoption
of water relative to other utility services varies with household income in developing countries.
Komives et al. (2003) have analyzed data on the percentage of households at different income
levels with four utility services: piped water, sewer, electricity, and telephone42 . As monthly
household income increases from very low levels to US$ 300 per month, coverage of all of
these infrastructure services increases rapidly; above US$ 300 coverage increases at a slower
rate. However, what is most striking is that, for households in this sample, at all income levels,
more people have electricity than piped water or sewer. Very few of the poorest households
have piped water or sewer, but almost a third of these households have electricity service. In
Kathmandu, Nepal, for example, all of the households surveyed had the option to connect to
electricity, water, and sewer; the majority of the very poor households chose electricity but
not water or sewer. As income increases, the percentage of households choosing water and
sewer increases, but the percentage with electricity is always higher. Almost no one, at any
income level, has only a piped water service; but many households have electricity and not
water. Thus, although most households would certainly like improved water and sanitation
services, it is not their most important development priority; given their limited resources,
many of them want electricity before an in-house piped water or sewer connection43 .
In short, the fact that water itself is a necessity does not necessarily mean that people
prefer piped water over electricity service. Indeed, because water is a necessity, households
must already have some access to water supply. The question is thus how much they value an
improved supply. This will depend on how bad the existing water service is, and how much
better the improved service is expected to be44 .

3.7 The heterogeneity of water


It is common to talk of the value of water as though it were a single, homogenous commodity.
This is obviously false: water has many dimensions besides just quantity. These include:
(a) location; (b) timing; (c) quality; and (d) variability/uncertainty. To a user, one liter of water
is not necessarily the same as another liter of water if it is available at a different location, at a
different point in time, with a different quality, or with a different probability of occurrence.

41
The subsequent experience when the Bank went ahead and implemented the projects has borne out
the predictions of the contingent valuation surveys quite well (Griffin et al., 1995). It would be wrong
to infer too much from the data on vended water purchases because the vast majority of households in
developing countries do not buy from vendors. This implies that their WTP for vended water is less than
the cost.
42
The data come from the World Bank’s Living Standards Measurement Surveys covering over 55,000
households in 15 developing countries.
43
This does not mean that water and electricity are not complements – they often are. It simply suggests
that people with limited budgets who cannot afford both generally prefer to have electricity first and
piped water later.
44
It may also depend on the quantity of water involved – because of diminishing marginal utility, what
a household would be willing to pay per unit for the first 20 L/d per capita is not necessarily the same
as it would pay to go from 40 to 60 L/d per capita.
The economic conception of water 81

There are two ways to incorporate the multi-faceted nature of water in a formal economic
analysis. The first approach is simply to define different types of water as different commod-
ities. For example, the consumption of water in January is represented by x1 , that in February
is represented by x2 , that in March by x3 , etc. The consumer is then assumed to have a
utility function defined over monthly consumption throughout the year and also over other
commodities whose consumption is denoted by z (which can be a vector or a scalar), leading
to the formulation:
u = u(x1 , x2 , . . . , x12 , z) (1)
The significance of this formulation is that it leads to separate demand functions for con-
sumption in each month. The demand for water in the ith month will be a function of the price
of water in that month, the prices of water in the other months (which may or may not be
different), and the price of z, as well as the consumer’s income, xi = hi (p1 , p2 , . . . , pN , pz , y).
The differences between one month’s demand function and that of another will reflect the dif-
ferent ways in which the two monthly consumptions enter the underlying utility function (1).
Given this approach, the annual demand for water is the aggregate of the 12 separate demand
functions for the individual months. There is no demand function for annual consumption
of water per se, except in the special case where the underlying utility function takes the
particular form:
 
u = u(x1 , x2 , . . . , x12 , z) = u xi , z (2)

This is the only formulation


 that generates a well-defined demand function for aggregate
annual consumption, X ≡ xi . Note, however, that the formulation in (2) implies that water
consumption in any month is a perfect substitute for consumption in any other month45 . More
generally, if one discounts the difference between facets of water use and treats water as a
single, homogeneous commodity, this is equivalent to assuming that the different types are
all perfect substitutes for one another. It is an empirical question whether this is a plausible
assumption.
The alternative framework for analyzing differentiated commodities was provided in a
somewhat simple form by Lancaster (1966) and then broadened by Maler (1974), and is
known as the characteristics approach to consumer demand46 . The Lancaster-Maler model
extends the utility model (1) by offering an explicit account of why the x’s are viewed as
separate commodities, based on their specific characteristics. The notion is that there is a set
of characteristics or attributes associated with each commodity. Suppose there are K relevant
characteristics (attributes), and let qik denote the amount or level of the kth characteristic asso-
ciated with one unit of consumption of commodity i. The characteristics of each commodity

45
Two commodities are said to be a perfect substitutes in consumption if the consumer is willing to trade-
off one for the other at the same, fixed rate of exchange regardless of how much or how little is consumed;
in his eyes they can always be used in exactly the same way, with exactly the same outcome. When two
commodities are perfect substitutes, they have essentially the same value. The polar opposite of perfect
substitute is perfect complement. Two commodities are perfect complements if they are valued in fixed
proportions to one another; consequently, they will always be purchased together in fixed proportions.
In this case, no value is placed on increasing one of the items unless there is a corresponding increase in
the other; an old-fashioned example (in England) was tea and milk.
46
The application of Maler’s work to the modeling of differentiated commodities was exposited in
Hanemann (1982).
82 W.M. Hanemann

are taken as given by the consumer who is free to vary only the quantity of the commodity, xi .
Thus, if the consumer wishes for more of the kth characteristic, she accomplishes this not by
changing the characteristics of any good, since these are fixed to her, but rather by switching her
consumption towards commodities with a high level of this characteristic (i.e. a high value of
qik ); quality variation is accomplished through quantity variation. If there are N separate differ-
entiated commodities together with undifferentiated consumption, z, the utility function takes
the form:
u = u(x1 , x2 , . . . , xN , q1 , q2 , . . . , qN , z) (3)

where qi ≡ (qi1 , . . . , qiK ). The demand functions for commodities now depend on the attributes
as well as the prices, and take the form xi = hi ( p1 , p2 , . . . , pN , q1 , q2 , . . . , qN , pz , y),
i = 1, . . . , N . Thus, this formulation provides a framework for analyzing the effect of
characteristics/attributes on demand – it provides a model of the demand for attributes
(i.e. for q).
I have focused so far on the multi-faceted nature of water with respect to consumer choice,
but this obviously applies to producer choice also. Both of the approaches described above
can be incorporated in a production function just as in a utility function. For example, the
production analog of (3) is a production function of the form:

y = f (x1 , x2 , . . . , xN , q1 , q2 , . . . , qN , z) (4)

where y is output, the x’s are forms of water input, and z is a vector of non-water inputs.
Suppose, for example, that y is crop production, N = 2, x1 is the quantity of groundwater
pumped by a farmer, and x2 is the quantity of surface water delivered to the farmer by the
irrigation district in which he is located. Even if the price were the same, the experience in
California has been that most farmers do not consider groundwater and surface water to be
equally attractive. They find groundwater more convenient because they totally control its
supply and can obtain it at the flick of a switch while, with surface water, they have to wait
until the irrigation district is able to route the water to them through the canal system. On the
other hand, in some parts of California there are differences in water quality, with groundwater
being more saline than surface water. Thus, immediacy of access and salinity are two of the
attributes that enter the crop production function (4) in this case.
In addition to providing a framework for conceptualizing the demand for q, whatever this
may be, the Lancaster-Maler model also provides a framework for the economic valuation of q.
It can thus be used to measure water users’ WTP for better availability of water, or less saline
water, or a more reliable water supply, or more generally water of one type versus another
(e.g. the premium on groundwater versus surface water, or on water at one location versus
another).
It should be emphasized that the attributes in q that differentiate one type of water from
another are by no means limited to the type of physical characteristics mentioned so far, such
as location, timing, quality, and reliability. Other aspects, such as how the water is provided,
can be the object of people’s concern and the focus of their preference. With water, users often
care greatly about fairness in allocation or payment – what is known as procedural justice –
and this may differentiate one source of water from another in their eyes. The Lancaster-Maler
formulation permits one to incorporate in q such psychological or sociological attitudes within
an economic model of the demand for water, so that one can analyze how these attitudes might
The economic conception of water 83

generate a different value for water when provided in a particular way or obtained from a
particular source47 .

3.8 The fallacy of using average value


In most policy-related applications of economic valuation involving water, the relevant quan-
tity that needs to be known is the marginal value of water rather than the average or total value.
Precisely because water is a necessity of life, most people have some access to some amount of
water, and most policy interventions therefore involve changing the quantity and/or quality of
access rather than transforming the situation from no access to some access48 . Ceteris paribus,
there is likely to be some degree of diminishing marginal utility for consumers, and dimin-
ishing returns for producers; for this reason there can be a substantial difference between the
marginal value of an increase in water supply and its average value. This needs to be empha-
sized because researchers often use an estimate of the average value of water to measure the
benefits of a policy intervention; the resulting estimate is likely to be inaccurate.
An example comes from the recent Spanish National Hydrological Plan, which proposed
a major water transfer from the Ebro River to the Mediterranean coast, from Barcelona
in the North to Murcia in the South (MIMAM, 2000). Of the 1050 Mm3 to be transferred,
560 Mm3 was targeted for delivery to agricultural areas along the coast that have been relying
on depleting supplies of groundwater. The correct way to measure the benefit from an incre-
ment in water supply for farming in the receiving areas is to estimate the marginal value of
water (marginal net profit) in the agricultural uses that would go out of production without
the importation of project water. Instead, the economic assessment performed by the Spanish
government (MIMAM, 2002) valued the imported water using an estimate of the average value
of water in current uses, calculated as the simple ratio of aggregate farming profit in the area
divided by aggregate water use. There are two flaws in this approach: 1) it interprets all profit
from farming in the area as exclusively a return to water; and 2) it treats the return to water as
constant regardless of the amount of water used.
Rather than just being a return to water, the profit from agriculture is likely to be a return
to the farmer’s investment in land and other fixed assets, and also a return to the farmer’s
own labor and his family’s labor. And, rather than the average value of water being constant
in the receiving areas, there are several reasons to believe that the average value declines as
more water is supplied, causing the marginal value of water to be less the average value. With
varying land quality and the opportunity to grow different crops, farmers in the region are
likely to respond to any reduction in water supply by idling their least productive land and
discontinuing their least profitable crops. Furthermore, some users of groundwater also have
access to some surface water supply – in varying ways and to varying degrees, the supplies
of water for farmers in the receiving area are interconnected so that, within the area, water is
a somewhat fungible commodity. Since the imported water is one among several sources of
water for irrigation, the relevant demand function is the farmers’ demand for project water,
not their demand for water overall. Because of the availability of substitutes, the agricultural
demand for project water is likely to be more elastic than the demand function for water overall.

47
What Frey & Stutzer (2005) refer to as procedural utility can be represented by this model.
48
By this, I am not implying that interventions are unimportant or trivial in their consequences.
84 W.M. Hanemann

Therefore, the marginal value of the imported water in the region is likely to be substantially
less than the existing average value49 .
The key difference between the two concepts is that the marginal value involves the deriva-
tive of a relationship, and to estimate this one needs a (formal or informal) model of how
water generates value. The average value, by contrast, can be estimated crudely by dividing
two quantities without any understanding of how they are related in reality, and without any
assurance that this ratio will remain constant. Consequently, the use of an estimate of average
value, and the assumption of its constancy, although common, are almost certainly a mistake.

3.9 The benefits of water


There are numerous ways in which an increment in access to water might produce benefits,
whether to those who use the water directly or to others. Examples include: the use of water
for agricultural or industrial production, its use for hydropower or for navigation, residential
use, flood control, water based recreation, or aquatic habitat. A key tool used by economists
in formalizing many of these benefits is the concept of a production function. A production
function is conceived as an empirical, causal relationship between the levels of inputs required
to produce an output, or an outcome, and the level of output or the outcome that results. One
example is the production function for an industrial or agricultural output as a function of water
and other inputs to the production process. Another example would be a health production
function relating inputs (including behavior patterns and levels of resource availability) to the
production (attainment) of health status outcomes. A third example would be an ecological
production function relating inputs and resource endowments to the production (attainment)
of ecosystem outcomes.
However, while the notion of a production function is undoubtedly useful as a conceptual
tool for organizing one’s thought about these matters, it may work less well as a dependable
empirical construct. It may work better on a micro-scale (i.e. at a factory level) rather than
at the level of an entire regional economy, and it implies a notion of causation that may be
oversimplified. In practice, it can often turn out to be surprisingly difficult to measure a
production function on a regional scale or, more generally, to measure the specific increment
in benefits associated with an increment in water availability; these difficulties are clearly
evident in the literature on water and economic development50 .
The notion that water supply contributes to economic growth and development seems
intuitively obvious. After all, it is known that many of the world’s major cities owe their ori-
gin to their location along coasts or rivers where water-borne transportation was facilitated.
But, the relevant question is whether an increment in water availability now would generate
an increment in economic activity now, and how much. In the USA, federal water projects

49
If one were considering the total elimination of farming across a large portion of the irrigated acreage in
the receiving areas, the difference between the average and marginal value could be of little significance,
but in this particular case the change involves about 10% of the irrigated farmland and 9% of the water
supply in the receiving area. An important piece of evidence suggesting that the NHP was substantially
overstating the value of the project water is the fact that the cost of irrigation water in the receiving areas,
including the cost of groundwater pumping, is mainly in the range of 0.06–0.27 a/m3 , while the NHP
valued the imported water at 0.75 a/m3 .
50
Similar difficulties are to be found in the literature on water and health.
The economic conception of water 85

have long been advocated for their claimed contribution to regional economic development.
However, the actual empirical evidence is less obvious and more negative. As Howe (1968)
noted, in industrial processes, water costs are a relatively small fraction of total production
costs even in water-intensive industries, and there are many examples of firms in such indus-
tries choosing to locate plants in water deficient areas because of market or non-water input
considerations.
In the USA in the late 1960s, there was a flurry of efforts to conduct formal, ex post statis-
tical analyses of the impact of water availability on economic development; the findings were
generally quite negative51 . The studies were motivated in part by Bower’s (1964) hypothesis
that the availability of water at the intake end and/or the effluent end is not a major factor
in macro-location decisions of industry relating to location in major geographical areas or
regions, such as river basins, but it can be a major determinant of micro-location decisions
relating to location within the region or basin. Ben-David (1966) found that employment
in the major water-intensive industries was not significantly related to a measure of water
availability in a cross-section regression of USA states, but there was a significant positive
impact at the county level in a regression of counties within Pennsylvania. Howe (1968)
extended this analysis to include all the counties in the USA, and found the evidence for such
an effect to be extremely limited. Cox et al. (1971) examined counties in the Northeastern
USA in which large water projects had been constructed between 1948 and 1958, and found
no relationship between project size and economic growth over the period 1950–1960. By
contrast, Garrison & Paulson (1972) examined counties in the Tennessee Valley region and
found a significant micro-location relation between water-oriented manufacturing employ-
ment and a measure of water availability. At the macro-location level, Carson et al. (1973)
sampled counties in geographic sub-regions from all parts of the country, both rural and urban,
and found no significant relationship between federal water resource projects and population
growth. Cicchetti et al. (1975) extended this study using economic sub-regions as the unit of
analysis and found that variables representing federal investment in irrigation facilities had
no significant impact on regional income and growth, and only a small and not convinc-
ingly significant impact on the value of farm output. There was some relationship between
economic growth and federal investments in flood control, hydropower and recreation, but
the coefficients were often unstable52 . A similar study by Fullerton et al. (1975) of coun-
ties in seven western states found no relationship between water investment and economic
growth.
It seems clear that an investment in water supply does not automatically guarantee economic
growth. But, what conclusion can be drawn? Is there never an economic case for investing in
water supply?
I want to suggest that part of the problem arises from the inadequate concept of causation
that is being utilized by economists in conceptualizing the notion of a production function.

51
The focus of these studies was long-run employment and economic growth after the completion of
project construction: it was recognized that there would be a short-term increase in employment during
the period of construction.
52
Some of these results and those of Cox et al. (1971) raise the question of the direction of causation: do
federal investments in flood control, say, cause economic growth, or do growing areas use their political
clout to attract federal water investments? Walker & Williams (1982) suggest that the latter is the causal
connection.
86 W.M. Hanemann

The philosophy literature makes a distinction between necessity and sufficiency: X could be
a necessary but not sufficient condition for Y to occur, or it could be a sufficient but not a
necessary condition. However, this distinction is generally ignored in the economic literature,
including both the theoretical and empirical analyses of the relationship between water and
growth. The production function as conventionally formulated as a relationship along the
lines of Y = f (X , Z, . . .), implies that X , Z and the other factors on the right-hand side of the
equation are each a sufficient condition for producing Y : changing any individual element X
or Z is sufficient to induce a change in the value of Y . Similarly, the conventional forms of
regression equation used in the statistical literature imply that the regressors are each sufficient
conditions for a change in the dependent variable.
However, the true relationship may be different, and perhaps more complicated. For one
thing, it seems plausible that having an adequate supply of water might be a necessary but not a
sufficient condition for economic growth. While water does not automatically generate growth,
it may be the case that areas which persist in lacking an adequate water supply (regardless
of whether or not they started out with adequate water) will not flourish economically. For
example, one can expect that people will eventually leave those areas and migrate to other
areas that do have an adequate water supply. Thus, lack of water could be a sufficient condition
for economic decline or, to put it another way, water may be a necessary but not sufficient for
economic growth. But, this is not a relationship that is captured in the existing formulations
of production functions and regression equations53 .
In fact, the relationship between water and growth might be even more complicated. It
may be that there are multiple possible causal pathways, such that while there is some causal
linkage between water and growth, the linkage is sufficiently imprecise and variable that water
is neither a necessary nor a sufficient condition for growth. In effect, there is sometimes a
causal linkage, but not always. If this is so, it would require a new formalism to express this
type of relationship.
The statistical literature suggests at least two possible methods for estimating a relationship
that is heterogeneous in the manner just suggested. Both allow for multiple possible relation-
ships for the determination of Y , with a mechanism that is at least partly stochastic determining
the specific relationship that applies at any particular instance; within this framework, one
of the possible relationships can involve Z as a causal determinant of Y (e.g. as a sufficient
condition), while in others Z is not a determinant of Y . One statistical approach is the finite
mixture model (McLachlan & Peel, 2000), in which it is assumed that different observations

53
One way to operationalize the concept of necessity is through what is known as a switching regression.
The general structure of a switching regression, framed around a variable Z, is:

f1 (Z, X1 ; β1 ) + ε1 if Z ≤ 0
Y=
f2 (Z, X2 ; β2 ) + ε2 if Z > 0

Here Y is the dependent variable, fi ( ) represents a possibly non-linear functional relation, X1 and
X2 are vectors of explanatory variables (which may or may not be different), and β1 and β2 are the
corresponding coefficient vectors which are to be estimated. In the present context, Z measures the level
of water supply and to represent it as essential one would have to impose the condition that f1 (Z, X1 ;
β1 ) + ε1 = 0 whenever Z ≤ 0.
The economic conception of water 87

within the data belong to different groups to each of which a separate statistical relationship
applies. Here one estimates both the underlying relationships and also the group membership
probabilities for each observation in the data. The other approach is model averaging, of which
Bayesian model averaging (Raftery et al., 1997; Hoeting et al., 1999) is receiving considerable
attention. This approach assumes that there are many possible models that could apply to the
given data set taken as a whole. One proceeds by estimating each model under consideration
using the entire data, and then averaging these models. In the Bayesian version, the weights
used for averaging are the posterior probabilities that each of the models is the correct one.
This approach produces confidence intervals for coefficient estimates that formally account
for model uncertainty in addition to estimation uncertainty.
The twin issues of model uncertainty and the need to reflect other causal relations besides
sufficiency deserve to receive greater attention in the empirical evaluation of the benefits
of water not only for economic development but also for health, fish habitat, and other
outcomes54 .

4 THE PROBLEM OF WATER FROM AN ECONOMIC PERSPECTIVE

It is commonly said that the problem of water is not one of economics but politics, not one of
physical shortage but governance. This is partly correct, but not entirely. The generic problem
of water is one of matching demand with supply, of ensuring that there is water of a suitable
quality at the right location and the right time, and at a cost that people can afford and are
willing to pay.
The difficulty in accomplishing this is partly institutional and certainly includes prob-
lems of governance. However, some of the problems of governance themselves have an
economic explanation. The omnipresence of fixed costs in surface water supply creates a
classic economic problem of cost allocation which has no satisfactory technical solution55 .
The extraordinary capital intensity and longevity of surface water supply infrastructure, and
the predominance of economies of scale, create a need for collective action in the provision
and financing of water supply that simply does not arise with most other commodities. It has
been recognized since Olson (1965) that the provision of goods through collective action may
be flawed because of a failure of incentives.
Olson set out to challenge the optimistic notion that individuals with common interests
can necessarily be counted on to act voluntarily to further those common interests. The prob-
lem arises from harmful coincidences of rivalness/non-rivalness in benefits combined with
excludability/non-excludability in costs. Examples are free riding by members of the group
who withhold their individual contribution but can still benefit from the results of their col-
leagues’efforts, and rent seeking by individuals who seek to capture for themselves the benefits

54
A non-water-related health application of Bayesian model averaging is Koop & Tole (2004). A recent
book by political scientists has been devoted entirely to the empirical modeling of necessary conditions
(Goertz & Starr, 2003).
55
The most convincing solutions are rooted in bargaining theory and identify a cost allocation based
on relative bargaining strength; this is more a political than an economic approach (Young, 1986). An
early application of bargaining theory to water is Rogers (1969); for a recent survey see Carraro et al.
(2005).
88 W.M. Hanemann

of collective action while throwing the cost on others56 . Consequently, as Olson concluded:
“unless the number of individuals is quite small, or unless there is coercion or some other spe-
cial device to make individuals act in their common interest, rational self-interested individuals
will not act to achieve their common or group interest”.
The challenge, thus, is to find a suitable non-coercive mechanism that motivates collective
action. This has become the subject of vast literature in economics, political science, sociology
and game theory. Success can be achieved, in principle, in several different ways. Cultural and
social norms shape preferences and may tilt the balance of individual choice in favor of col-
lective action. Homogeneity can help in some circumstances57 . The nature of the institutional
arrangements is crucial58 : if the rules are simple, transparent and devised locally, if moni-
toring and enforcement are relatively cheap, with graduated sanctions for non-compliance, if
low-cost and fair adjudication is available, then, ceteris paribus, successful collective action is
more likely. However, the extent to which these conditions are met depends partly on people’s
outlook and disposition, and partly on the physical reality. For the reasons enumerated above,
water supply and sanitation are not always well situated in this regard.
In short, while there clearly are some distinctive emotive and symbolic features of water that
make the demand for water different, there are also some distinctive physical and economic
features which make the supply of water different and more complex than that of most other
goods. This fact has often been overlooked by economists and non-economists alike.

REFERENCES

Agrawal, A. (2002). Common Resources and Institutional Responsibility. In: E. Ostrom; T. Dietz;
N. Dolsak; P.C. Stern; S. Stonich & E.U. Weber (eds.), The Drama of the Commons. National Research
Council.
Anderson, L.D. (1980). The diffusion of technology in the 19th century American City: municipal water
supply investments. Ph. D. Dissertation. Northwestern University, Chicago, USA.
Armstrong, M.; Cowan, S. & Vickers, J. (1994). Economic analysis and British experience. MIT Press.
Baland, J.M. & Platteau, J.P. (1999). The ambiguous impact of inequality on local resource management.
World Development, 27(5): 773–788.

56
Spiller & Savedoff (1999) emphasize the possibility of rent seeking by the government. They
argue the combination of economies of scale and massive sunk costs of investment encourage gov-
ernments in developing countries to act opportunistically against the private (and often foreign)
companies that operate the water supply system by pressuring them to lower prices, disallowing their
costs, requiring them to undertake special investments, controlling their purchasing or employment
patterns, or trying to restrict the movement or composition of capital. “All these are attempts by
politicians . . . to capture the rents associated with the company’s sunk costs by administrative mea-
sures”. Spiller & Savedoff argue that this leads to a low-level equilibrium of low prices and bad
service.
57
The ambiguous implications of inequality for the success of collective action are noted by Baland &
Platteau (1999) and by Bardhan & Dayton-Johnson (2002). A disequalizing shift in the access to common
resources has two effects which work in opposite directions. On the one hand, the people who benefit
from the change have a larger stake in the resource and, therefore, a greater incentive to act towards the
collective good. But, there is a corresponding disincentive to the other individuals whose share of the
outcome has been reduced.
58
This has been emphasized most influentially by Elinor Ostrom; see, for example, Ostrom (1990, 2003).
I draw here on Agrawal (2002) who provides a useful summary of the literature.
The economic conception of water 89

Bardhan, P. & Dayton-Johnson, J. (2002). Unequal irrigators: heterogeneity and commons management
in large-scale multivariate research. In: E. Ostrom; T. Dietz; N. Dolsak; P.C. Stern; S. Stonich &
E.U. Weber (eds.), The Drama of the Commons. National Research Council.
Barlow, M. & Clarke, T. (2002). Blue Gold: the fight to stop corporate theft of the World’s Water. The
New Press.
Baumann, D.D. & Boland, J.J. (1998). The case for managing urban water. In: D.D. Bauman; J.J. Boland
& M. Hanemann (eds.), Urban water demand management and planning. McGraw Hill: 1–30.
Ben-David, S. (1966). Impact of water resource development on economic growth. In: Quarterly Progress
Report to the Corps of Engineers, July. Raleigh, North Carolina, USA.
Blake, N.M. (1956). Water for the cities. Syracuse University Press.
Bowen, H.R. (1943). The interpretation of voting in the allocation of economic resources. Quarterly
Journal of Economics, 58(4), November: 27–48.
Bower, B.T. (1964). The location decision of industry and its relationship to water. Paper presented at
the Conference of the Committee on the Economics of Water Resources Development. December, San
Francisco, California, USA.
Bowley, M. (1973). Studies in the History of Economic Theory before 1870. Macmillan.
Carraro, C.; Marchiori, C. & Sgobbi, A. (2005). Application of negotiation theory to water issues.
Fondazione Eni Enrico Mattei. Working Paper 65, May.
Carson, J.M.; Rivkin, G.W. & Rivkin, D.M. (1973). Community Growth and Water Resources Policy.
Praeger.
Cicchetti, C.J.; Smith, V.K. & Carson, J. (1975). An Economic Analysis of Water Resource Investments
and Regional Economic Growth. Water Resources Research, 11(1), February: 1– 6.
Ciriacy-Wantrup, S.V. (1947). Capital returns from soil-conservation practices. Journal of Farm
Economics, 29: 1181–1196.
Clawson, M. (1959). Methods of measuring the demand for and value of outdoor recreation. Reprint
no. 10. Resources for the Future, Inc., Washington, USA.
Cox, P.T.; Grover, W.W. & Siskin, B. (1971). Effect of water resource investment on economic growth.
Water Resources Research, February: 32–38.
Davis, R.K. (1963). Recreation planning as an economic problem. Natural Resources Journal, 3(2):
239–249.
Dupuit, J. (1844). De la Mesure de l’Utilite des Travaux Publiques. Annales des Ponts et Chaussees, 2nd
Series, 8. Reprinted in translation (1952) as: “On the Measurement of the Utility of Public Works”,
in: International Economic Papers, 2: 83–110.
Erie, S.P. & Joassart-Marcelli, P. (2000). Unraveling Southern California’s Water/Growth Nexus:
Metropolitan Water District Policies and Subsidies for Suburban Development, 1928–1996. California
Western Law Review, 36, Spring: 267–288.
Frey, B.S. & Stutzer, A. (2005). Beyond outcomes: measuring procedural utility. Oxford Economic
Papers, 57(1), January: 90–111.
Fullerton, H.; Lewis, W.C.; Anderson, J.C.; Keith, J.E. & Willis, R. (1975). Regional development: an
econometric study of the role of water development in effectuating population and income changes.
Report PRRBEO89–1. Utah Water Research Laboratory, Utah State University, Logan, USA.
Garrison, C.B. & Paulson, A.S. (1972). Effect of water availability on manufacturing employment in the
Tennessee Valley region. Water Resources Research, April: 301–307.
General Accounting Office (1996). Information on allocation and repayment of costs of constructing
water projects. Bureau of Reclamation (GAO/RCED-96-109).
Gleick, P.H. (1999). The Human Right to Water. Water Policy, 1(5): 487–503.
Glennon, R. (2004). The Price of Water. Journal of Land, Resources & Environmental Law, 24(2):
337–342.
Goertz, G. & Starr, H. (2003). Necessary Conditions. Rowman & Littlefield.
Gordon, B. (1975). Economic analysis before Adam Smith. Macmillan Press.
Griffin, C.; Briscoe, J.; Singh, B.; Ramsasubban, R. & Bhatia, R. (1995). Contingent Valuation and
actual behavior: predicting connections to new water systems in the State of Kerala, India. The World
Bank Economic Review, 9: 373–395.
Hamlin, C. (2000). Waters or Water? – Master Narratives in Water History and their implications for
Contemporary Water Policy. Water Policy, 2: 313–325.
90 W.M. Hanemann

Hammond, R.J. (1960). Benefit-Cost analysis and water pollution control. Food Research Institute,
Stanford University Press.
Hanemann, W.M. (1982). Quality and Demand Analysis. In: G.C. Rausser (ed.), New Directions in
Econometric Modeling and Forecasting in U.S. Agriculture. Elsevier/North Holland, Inc., New York,
USA: 55–98.
Hanemann, W.M. (1991). Willingness to Pay and Willingness to Accept: how much can they differ?
American Economic Review, 81(3): 635–647.
Hanemann, W.M. (2002). The Central Arizona Project. Department of Agricultural and Resource
Economics. Working Paper 937, October. University of California, Berkeley, USA.
Henderson, A. (1941). Consumer’s surplus and the compensating variation. Review of Economic Studies,
8: 117–121.
Hicks, J.R. (1939). Value and Capital. Oxford University Press.
Hicks, J.R. (1941). The rehabilitation of consumer’s surplus. Review of Economic Studies, 8: 108–116.
Hicks, J.R. (1942). Consumer’s surplus and index numbers. Review of Economic Studies, 9: 126–137.
Hicks, J.R. (1943). The four consumer’s surpluses. Review of Economic Studies, 11: 131–141.
Hicks, J.R. (1946). The generalized consumer’s surpluses. Review of Economic Studies, 15: 27–33.
Hoeting, J.A.; Madigan, D.; Raftery, A.E. & Volinsky, C.T. (1999). Bayesian model averaging: a tutorial.
Statistical Science, 14(2): 382–417.
Howe, C.W. (1968). Water resources and regional economic growth in the United States, 1950–1960.
Southern Economic Journal, 34(4), April: 477–489.
Hurwicz, L. & Uzawa, H. (1971). On the integrability of demand functions. In: J.S. Chipman;
L. Hurwicz; M.K. Richter & H.F. Sonnenschein (eds.), Preferences, utility and demand. Harcourt
Brace Jovanovich, New York, USA: 114–148.
Komives, K.; Whittington, D. & Wu, X. (2003). Infrastructure coverage and the poor: a global perspective.
In: P.J. Brook & T.C. Irwin (eds.), Infrastructure for poor people: public policy for private provision.
The World Bank: 77–123.
Koop, G. & Tole, L. (2004). Measuring the health effects of air pollution: to what extent can we really
say that people are dying from bad air? Journal of Environmental Economics and Management, 47:
30–54.
Krutilla, J.V. (1967). Conservation reconsidered. American Economic Review, 57: 787–796.
Kupel, D.E. (2003). Fuel for growth: water and Arizona’s urban environment. University of Arizona
Press, USA.
Lancaster, K. (1966). A new approach to consumer theory. Journal of Political Economy, 74:
132–157.
Lockwood, M. (1996). Non-compensatory preference structures in non-market valuation of natural area
policy. Australian Journal of Agricultural Economics, 40(2): 85–101.
Maler, K.G. (1971). A method of estimating social benefits from pollution control. Swedish Journal of
Economics, 73: 121–133.
Maler, K.G. (1974). Environmental Economics: a theoretical enquiry. Johns Hopkins University Press
for Resources for the Future.
Marshall, A. (1879). The pure theory of (domestic) values. London School of Economics, UK.
Marshall, A. (1890). Principles of Economics. 1st edition. Macmillan, London, UK.
McLachlan, G. & Peel, D. (2000). Finite mixture models. John Wiley.
MIMAM (Ministerio de Medio Ambiente [Ministry of Environment of Spain]) (2000). Plan Hidrológico
Nacional: análisis económicos. September. Madrid, Spain.
MIMAM (Ministerio de Medio Ambiente [Ministry of Environment of Spain]) (2002). Strategic
assessment of the National Hydrological Plan. Summary Document. January. Madrid, Spain.
Niehans, J. (1990). A History of Economic Theory: classic contributions, 1720–1980. Johns Hopkins
University Press.
Olson, M. (1965). The Logic of Collective Action: Public Goods and the Theory of Groups. Harvard
University Press.
Ostrom, E. (1990). Governing the Commons. Cambridge University Press.
Ostrom, E. (2003). How types of goods and property rights jointly affect collective action. Journal of
Theoretical Politics, 15(3): 239–270.
The economic conception of water 91

Postel, S. & Vickers, A. (2004). Boosting Water Productivity. In: Worldwatch Institute, State of the World
2004. W.W. Norton.
Pribram, K.H. (1983). A History of Economic Reasoning. Johns Hopkins University Press.
Raftery, A.E.; Madigan, D. & Hoeting, J.A. (1997). Bayesian model averaging for linear regression
models. Journal of the American Statistical Association, 92, March: 179–191.
Ridker, R.G. (1967). Economic costs of air pollution. Praeger.
Rogers, P. (1969). A Game Theory Approach to the Problems of International River Basins. Water
Resources Research, 5(4), August.
Samuelson, P.A. (1954). The pure theory of public expenditure. Review of Economics and Statistics,
36(4), November: 387–389.
Schumpeter, J.A. (1954). History of Economic Analysis. Oxford University Press.
Shiva, V. (2002). Water wars: privatization, pollution, and profit. South End Press.
Spiller, P.T. & Savedoff, W.D. (1999). Government opportunism and the provision of water. In: P.T. Spiller
& W.D. Savedoff (eds.), Spilled water: institutional water. Institutional Commitment in the Provision
of Water Services. Inter-American Development Bank. Washington, USA.
Tarr, J.A. (1979). The separate vs. combined sewer problem: a case study in urban technology design
choice. Journal of Urban History, 5, May.
Trice, A.H. & Wood, S.E. (1958). Measurement of recreation benefits. Land Economics, XXXIV. August.
UNESCO (2000). World Day for Water 2000. See: http://www.unesco.org/science/waterday2000/
Variability.htm
Walker, R.A. & Williams, M.J. (1982). Water from power: water supply and regional growth in the Santa
Clara Valley. Economic Geography, April: 95–119.
Ward, F.A. & Michelsen, A. (2002). The economic value of water in Agriculture: concepts and policy
applications. Water Policy, 4: 423–446.
Wegge, T.C.; Hanemann, W.M. & Loomis, J. (1996). Comparing benefits and costs of water resource
allocation policies for California’s Mono Basin. Economic institutions and increasing water scarcity.
In: D.C. Hall (ed.), Advances in the economics of environmental resources. Volume 1: Marginal cost
rate design and wholesale water markets. JAI Press, Greenwich, Connecticut, USA: 11–30.
Weisbrod, B.A. (1964). Collective consumption services of individual-consumption goods. Quarterly
Journal of Economics, 78(3): 471–477.
Willig, R.D. (1976). Consumer’s surplus without apology. American Economic Review, 66: 589–597.
Wittfogel, K.A. (1957). Oriental despotism: a comparative study of total power. Random House Inc.
World Bank (1975). Issues in Village Water Supply. Washington, D.C., USA.
Worster, D. (1986). Rivers of Empire: water, aridity, and the growth of the American West. Random
House.
Young, H.P. (ed.) (1986). Cost allocation: methods, principles, applications. Elsevier Science Publishers.
Young, R.A. & Haveman, R.H. (1985). Economics of water resources: a survey. In: A.V. Kneese &
J.L. Sweeney (eds.), Handbook of Natural Resource Energy Economics, Vol. II. Elsevier Science
Publishers: 465–529.
CHAPTER 5

The value of water and theories of economic growth

M.S. Aguirre
The Catholic University of America, Washington, D.C., USA

ABSTRACT: Michael Hanemann (this volume) addresses important theoretical and empirical issues
that are relevant in ensuring an accurate economic analysis of market and non-market valuations of water,
its role in economic growth, and water management and policy issues. This short essay provides some
comments on the approach taken by Hanemann. These comments underline, among other issues, the
role of water use as well as the need to determine the price and cost of water, the need for intertemporal
analysis, the appropriateness of using marginal versus average values, and the importance of taking into
account institutional as well as water access conditions. All of these issues suggest that population has
a solution role in resolving some of the water management problems that we face today.

Keywords: Water value, economic growth theories, population and resources.

1 INTRODUCTION

At the heart of the water crisis debate lies one question: is there enough water to sustain
the present and the future population? How you answer this question depends greatly on
whether or not one sees population as a problem. Some would argue that population is a
problem in that resources (in this case water) are limited, that they can only sustain a certain
number of people (although no one knows what that particular number may be), that the
more numerous we become, the less resources each person will have available and it will
become more difficult to provide the resources that each consumer desires1 . Others argue that
population is not a problem. They contend that numbers in themselves do not equal scarcity;
rather, poorly structured societies as well as bad economic policies and governance bring about
scarcity and poverty2 . How people perceive the issue of population is critical, for it is by these
perceptions that international legislative policies are formulated, and local water projects
and policies are designed and approved. Thus, it is equally critical that people ensure that

1
Brown et al. (1999) capture the typical arguments used in this position. A further corollary of this
perspective, some times is identified as Neo-Malthusians. They see people as destroyers of resources and
violators of environmental limits. Among the leading representatives of this group are Ehrlich and Hardin.
A more recent presentation of their argument can be found in Ehrlich & Ehrlich (1990), and Hardin (1998).
For a detailed analysis of their views see Simon (1996a, 1996b), Furedi (1997), and Wolfgram (1999).
2
See Sen (1981, 1994), Fukuyama (1999), and Kliksberg (2000, 2001). There are also those who attribute
the present problems not to population but to the distribution of resources given the present structures.
Some of these authors include Dobson et al. (1997), Matson et al. (1997), Rabkin (1997), Kiester (1999),
and Johnson (2000). It is clear that the conclusion of the authors in this volume is that the so call crisis
of water is a crisis of governance, not of water availability.
94 M.S. Aguirre

their perceptions are grounded not in rhetoric and emotions, but in established scientific and
empirical data.
Michael Hanemann (this volume) addresses important theoretical and empirical issues that
are relevant in ensuring an accurate economic analysis of market and non-market valuations
of water, its role in economy growth, and water management and policy issues. From this
perspective, the issues he raises significantly contribute to helping differentiate what is rhetoric
and emotions from the real issues in the so-called water crisis debate.
This short essay will provide some comments on the approach taken by Hanemann in
addressing the issues raised in his chapter, and why they are relevant for economic growth
theory.

2 SOME COMMENTS

In analyzing the value of water and its relation to the water and diamond paradox, Michael
Hanemann, recalling Adams Smith and the Classics, notes the difference that exists between
the “value in use and the price (value in exchange)”, and underlines the important role that
marginal analysis plays in the completion of the discussion of the value of a good. This is so
because “it highlights the distinction between value and cost, between demand and supply”.
He thus accurately concludes, that “[t]he distinction between marginal and total [value] is the
key to the full resolution of the diamond and water paradox: water may have a smaller value
than diamonds at the margin, but it undoubtedly has a larger total value”. He then goes on
to address the valuation of non-market goods and emphasizes that the “use of non-market
valuation applies to positive as well as negative environmental impacts of water projects”.
Yet, while he acknowledges the shift in the argument regarding what determines the source
of value from “its inherent usefulness and ability to please man according to rules of reason”
to “subjective human preference rather than objective need”, Hanemann fails to note the
importance that such a shift can have when addressing both the market price and non-market
valuation of water. He also fails to acknowledge the importance of the externalities derived
from water, such as environmental restoration or deterioration. If subjective human preferences
are considered without taking objective human needs into account, one can run the risk, and
people often do, of using resources inefficiently through waste and/or misallocations. This in
turn leads to bad policy and poor water management. The issue becomes even more relevant
when this analysis is framed in an inter-temporal analysis, this being within a strictly economic
analysis context or that of city planners.
Addressing the essentialness of water, Hanemann accurately acknowledges that water “fits
the definition” of an essential commodity. It does so as an input of production for “agriculture
and also in several manufacturing industries”. It also does so as a final good because “human
life is not possible without access to 5 to 10 L/d of water per person”. He underlines the fact
that water, “in addition to being essential for human life, [it] contributes in important ways
to the enjoyment of the satisfaction of life”. From here, he draws two conclusions: (1) “the
fact that water is essential for human life is almost certainly irrelevant when assessing the
value of residential water supply [because the] … ways in which water is used are nowhere
near … level[s] at which essentialness applies”; and (2) “developing countries might also move
along a rising trajectory of residential water consumption because … life styles in developed
countries could also become attractive to people in these countries as their income rises”.
The value of water and theories of economic growth 95

Furthermore, he argues that “while it is obviously appropriate to think in terms of human


needs for water, it is also appropriate to recognize that people have demands for water as a
commodity that generates pleasure by utilizing it in various ways”. Therefore, he concludes,
“planners will need to adopt a behavioral approach to the analysis and projection of urban
demand, as opposed to the engineering/public health approach that dominates the literature on
water and poverty today”. Along these lines, Hanemann points to the fact that the question is
“how much [the household] values improved [water] supply” and not whether the “household
values access to water versus no access to water at all”. Since water is an essential good,
everyone has some way to access it. Otherwise, they would not survive3 .
While we agree to the benefits of a behavioral approach, we find the analysis of the role of
water essentialness in the assessment of the residential water supply’s value, only acceptable
if no intertemporal analysis is taken into account. Such an assumption, however, is certainly
mistaken in the context of water policy design and management. In an intertemporal analysis of
water supply, need becomes a very relevant factor, not only because the basic consumption
of present and future generations must be met, but also because it affects the optimum point
of present and future consumption as well as pricing. To ensure the access to clean water of
future generations, costs are involved and thus, necessity and use, cost and value, are not
and should not be disconnected (Asano, this volume, clearly speaks to this point). From this
point of view, one can also appreciate why access to not just improved water supply is relevant,
especially when considering this issue within the context of developing countries. Furthermore,
it speaks to the fact that, in some policy-related applications of economic valuation of water,
the marginal and average values need to be taken into account. This is true especially, but not
exclusively, in developing countries. Many households in developing countries might have
access to some water, but that access might not be humanly acceptable from an economic and
health point of view (for further comments on this point, see: Bergkamp, this volume; Sullivan,
this volume). This is a case in point where the “heterogeneity of water”, which Hanemann
accurately addresses in his chapter, clearly applies. Not all water is the same.
Finally, Michael Hanemann raises an issue that is very relevant in the context of growth
when he states that “water could be a necessary but not a sufficient condition for sustained
economic growth” and recommends the use of a “discrete mixture approach” for handling
the “causal, but not always” direct role that water has in real economic growth. As has been
previously mentioned, institutions and political stability as well as human, social, and moral
capital, play an important role in economic development and growth.

3 ECONOMIC THEORIES OF GROWTH

Based on what constitutes the main determinant of real economic growth, one can roughly
divide the economic theories of growth into four groups. Two of them have a negative

3
When addressing this issue, Hanemann recalls the case studied by Komives et al. (2003) and he
concludes that there is a preference of electricity over improved water sources in developing coun-
tries. He fails to acknowledge, however, that these are not necessarily exclusive of each other but can
be complementary. For example, access to electricity could allow consumers to access groundwater
through electric water pumps rather than piped water. The first one often is significantly cheaper than
the second one.
96 M.S. Aguirre

perception of population while the other two have a positive view. All schools often resort to
economics and environmental sciences to validate their theory.
The argument to support a positive view of population in the water-population-environment
relationship is mainly centered in the concepts of human, social, and moral capital, and they
find validation in empirical data (for a review of the literature, see: Simon, 1996a, 1996b;
Wolfgram, 1999; Eberstadt, 2000). On the other hand, the steps in the argument used to
support a negative view of population in the water-population-environment relationship can
be reduced to mainly three. (1) Rapid growth on population means the spread of scarcity
(less water available for each person or to satisfied the subjective human preferences for
water), and it is a main obstacle to economic growth in poor countries (as water is essential
for the process of growth), because it reduces or cancels potential improvements in living
standards and aggravates conditions such as poor health, sanitation, and malnutrition. (2) The
political implications of such trends threaten government stability in developing countries,
and encourage the confrontation between developed and developing countries. And (3) it
pushes future generations to scarcity, and an unsustainable environment (water) carrying
capacity. However, literature from both disciplines suggests opposite conclusions (Simon,
1996a, 1996b; Wolfgram, 1999; Eberstadt, 2000).

3.1 Classical Economic Growth: Malthusian Theory (this section relays


heavily on Aguirre, 2002)
Thomas Robert Malthus proposed in his 1798 Essay in the Principle of Population a relation-
ship between population growth and what he termed subsistence. The first grew geometrically
while the second increased only at an arithmetic ratio. Consequently, he claimed the existence
of an inverse relationship between population growth and economic development derived
from the law of diminishing returns4 . Malthus’ problem was that he failed to explore his
theory against historical experience; no theory can be said to be scientifically proven if that
theory cannot be verified by empirical evidence.
Malthus’ inverse relationship between population and growth leads Classical Economic
Theory to the following arguments: (a) The consumption effect (for a given amount of water,
population growth affects consumption directly); (b) The production effect on private and
public goods (population growth affects consumption indirectly through the effect on produc-
tion per worker. With fixed capital, average production per worker will suffer diminishing
returns. Along the same lines, with a fixed level of revenue, a larger population will increase
the demand for public services – utility services, education and health care – thus reducing
the quality of these services and indirectly hindering development through the reduction of
funds allocated to infrastructure); (c) Age-Distribution effect (a faster-growing population
implies a larger proportion of children and, given the amount of resources, a smaller output
per capita); and (d) Dilution of Capital (with a fixed income, population growth reduces
savings and human capital – education per person – and therefore reduces physical and human
investment.

4
This law implies that more people mean fewer goods (including water) for each person; thus, as
population grows, scarcity of water inevitably increases. He believed that man’s ability to increase his
food (including water) supply was constrained in three particular ways: through land (water) scarcity,
limit productive capacity of cultivated land, and the law of diminishing returns.
The value of water and theories of economic growth 97

In summary, and applying it to water, the classical theory of population growth, assuming
a fixed level of water, predicts a decrease in per capita water availability (and therefore
less economic growth or income) in two ways: more consumers divide any given amount of
water, and each worker produces less because there is less capital, private and public, per
worker. In addition, the growing number of young children poses an additional burden in the
reduction of consumption because they consume but they do not produce. Finally, population
growth hinders economic growth because, by reducing savings and water services, it reduces
investment in water technology. The key made in this theory is the ceteris paribus condition
(other things being equal) where resources (including water) are given and therefore constant.
However, when challenged, this theory fails both theoretically and empirically. When
applied to the case of water, this failure is also present. Analyses at both levels suggest that there
is no statistically proven simple relationship between population growth and economic growth,
population size and economic growth, population size and resources, or population growth and
environment5 . The absence of a correlation contradicts the conventional Malthusian deductive
conclusion. The only persuasive argument in the face of this absence of correlation, as Simon
(1996b) states, is a plausible scenario in which one or more specified variables that have been
omitted from the analysis would, in fact, lead to a negative relationship between population
growth and economic growth. Some of these factors can include poor water governance, new
technologies developed or those raised by J.A. Allan in the political arena. In this context,
Hanemann’s assessment on the type of modeling that should be used for water analysis seems
to be especially relevant.
In fact, since the seminal work of Coale & Hoover (1958), several studies have followed
supporting or contradicting the negative population view. The 2003 World Water Assessment
Programme concluded that “this crisis is one of water governance … caused by the ways in
which we mismanage water”. On the other hand, we do encounter problems of scarcity around
the world, problems of poor consumer selection (water waste), or utility services’design such as
the one mentioned by Hanemann regarding water and electricity (Llamas & Martínez-Santos,
this volume, address this issue).

3.2 Neo-Classical theory


The Neo-Classical model of growth, as presented by Solow (1956) as well as some models
of technological diffusion focuses on economic growth through investment, ignoring any link
between population and the economy. That is, adjustments in growth take place due to the
behavior of investment in physical capital (water technology and infrastructure in our case). In
these models, growth is a worldwide process and country characteristics determine the relative
level of income. Thus, low persistence is consistent with shocks of any size. Shocks may only
play a minor role in determining the long-run path of output, despite being an important
determinant of variance in decade-long growth rates.
Although these models have been able to explain the experience of developed countries
in certain cases, they have failed to explain the worldwide experience. Dorwick & Nguyen
(1989) present results for OECD countries that support the Neo-Classical theory. One may be
tempted to think also on Singapore, South Korea, and other Asian countries as well. Yet, in the

5
These works include Denison (1985), Rosemberg & Birdzell (1986), Scully (1988), Barro (1989),
Simon (1992, 1996a), Birdsall (1995), Eberstadt (1995), Agenor & Montiel (1999), Agenor (2000).
98 M.S. Aguirre

case of the first two countries mentioned, human as well as physical capital investment took
place as Blackburn & Ravn (1993) suggest.
From the point of view of this theory, achieving efficiency in water use and recycling through
new technologies and infrastructure is the key to economic growth and water sustainability.

3.3 Human capital theory


Gary Becker advanced a model that relates the concept of human capital to economic growth.
In doing so, he proposes an alternative to Malthusian models. In this theory, human capital
is introduced as an important source of economic development that depends on advances in
technological and scientific knowledge. A key assumption of this model is that the rate of
return on investments in human capital rises rather than declines as the stock of human capital
increases; man is creative and therefore the education of today implies more production in the
future. For this reason, resources are not necessarily fixed and may increase as the population
increases. Furthermore, Becker et al. (1993) found that population growth, when studied in
the light of human capital theory, leads to multiple equilibrium points: an underdeveloped
steady state with high birth rates and low levels of human capital, and a developed steady state
with low fertility and high stocks of human and physical capital. They concluded that this
means that history and luck are also critical determinants of a country’s growth experience6 .
This implies that population growth is not the only determining factor in economic devel-
opment, as the Malthusian theory claims. Training and educational programs together with
physical capital investment are the important factors though. But what about diminishing
returns? Becker found the answer to this issue in the increase of labor productivity due to
education, consequently rejecting the Malthusian assumption of fixed resources. Following
the concept of human capital, recent works have proposed models that relate population to
growth7 . They set forth an alternative to the Malthusian and Neo-Classical models of eco-
nomic growth by introducing human capital as an important source of economic development,
a source that depends on both technological and scientific knowledge8 .
Economic development has not been solely explained by the expansion of physical capital
per worker, as the Neo-Classical school has proposed, or by the decrease in population as
Malthus suggested. The human capital approach criticizes both perspectives because they
underestimate the dignity and creativity of the economic agent, thereby failing to acknowledge
this economic agent as the ultimate resource. Other issues, such as terms of trade, service of
the debt, the cost of intermediate goods, and institutional features of each country, including
political stability, are important for economic growth as well. Yet, it has been the introduction
of human capital that has shed new light on the understanding of the development process. It
is worth noting that the evolution of the discussion around economic development and water
supply also reflects this evolution. During the 1970s, the focus was on infrastructure, and in

6
Concerning the issue of luck, history and growth see Barro & Lee (1993), De Long & Summers (1993),
and Easterly et al. (1993).
7
Some of these works are analyzed within an overlapping generation model. See: Barro (1974), Becker
(1974, 1991), Razin & Ben-Zion (1975), Willis (1985), Becker & Barro (1988), and Becker et al. (1993).
King (1993) includes the proceedings of a conference on population and economic growth sponsored
by the World Bank.
8
Such findings have been long sustained by Julian Simon, Norman Macrae, Aaron Wildavsky, Ben
Wattenberg, Karl Zinsmeister, and others.
The value of water and theories of economic growth 99

the 1980s it was on economic liberalization. It was not until the 1990s when the focus shifted
towards institutions and water governance.
Within this view, one can appreciate the relevance of the distinction made by Hanemann of
“water as a necessary versus a sufficient condition for economic growth”. It also speaks of the
importance of incorporating intertemporal analysis in the determination of both the demand
and the supply of water, as has been previously mentioned.

3.4 Neo-Malthusian theory


Since the Agenda 21 of Rio (UNEP, 1992), the Malthusian theory has once more gained an
audience in the population debate. Supporters of this theory claim that population growth
is unsustainable not only with regard to food, but also with regard to resources such as oil,
minerals, land, and water. In 1968, two influential Neo-Malthusian works, Ehrlich’s Population
Bomb (1968) and Garrett Hardin’s Tragedy of the Commons (1968), issued warnings about
the limits of sustenance and of resources, which they claimed were doomed. Within this
theoretical framework, there are two main sub-categories: The Limited Resource Perspective
and the Socio-Biological Perspective. The former takes the classic Malthusian argument and
applies it to all natural resources, while the latter, almost acting as a sub-set of the former,
treats the environment as a limited resource and regards people as a threat to the biodiversity
and ecological balance of that resource.
Both perspectives have failed to produce sound projections because they lack sound data and
logic. For example, Paul Ehrlich claimed in 1968 that “hundreds of millions” of people would
die of starvation by the 1970s, that 65 million Americans would starve, that the population
of the USA would decline by 22.6 million persons, and that England would cease to exist
by 2000 (Ehrlich, 1968). More recently, Ehrlich & Ehrlich (1990) renewed these predictions,
although with more caveats, since the original predictions failed to materialize. Statistical data
regarding water provided by international bodies such as the United Nations and the World
Bank as well as by the experts’ evidence presented in this volume, however, fail to support their
position. The world is nowhere near the mass starvation or water crisis predicted by Ehrlich
or Brown, yet much policy has been designed based on their assumptions not only in the area
of water management but on other resources as well.
Supporters of market forces (Fullerton & Stavins, 1998, among others), argue that the
market will correct for inefficiencies, and that carefully constructed initiatives can help to
guide the market, particularly in the area of environmental-water protection. There are also
those who disagree with this perspective because they attribute the present problems not to
population but to the distribution of resources given the present structures9 . What we know
is that among poor countries there are some that have a high rate of population growth and
others have too little population. The overall decrease in population growth (the number of
children per women worldwide has decreased from 3.6 in 1980–1985 to 2.7 in the present)
has not helped some countries to overcome poverty. The population has not become poorer in
spite of having increased, but it has produced beyond the subsistence level. The productivity
of the earth’s land and water availability has grown more quickly than the world’s population.
In addition, the quality of life has increased also in less developed countries according to the
2003 Human Development Report.

9
See footnote 2 for references.
100 M.S. Aguirre

It is within this theory where, I believe, the importance of addressing the impact of the
shift from objective need to subjective human preference in determining the value of water is
made more obvious, whether it is analyzed at the margin (where it should be analyzed in most
cases) or at the level (total value which should be use when the quality of the access to water
is being considered). When subjective human preferences (level of consumption, comfort,
or expectations) are stressed over the objective human need, economic and supply-planning
water analysis can mislead policy decisions. A clear case is the wasted water seen in developed
countries: an average of 200 to 600 L/d of water consumption versus 47 L/d in developing
countries (UN Statistics Division, 2003, http://unstats.un.org/unsd/environment/), or the use
of leaks in water pipes for the provision of water at a higher cost, leading the poorest to high
price differentials. Such situations speak to the importance of addressing the need to focus
not only on the management of the supply, but also of the demand. A good way to begin to
address this problem, as Llamas & Martínez-Santos (this volume) point out, is the elimination
of perverse subsidies, which disturb the market price.

4 CONCLUSION

The economic value of any good, including water, has experienced an evolution accompanied
by both theoretical and empirical tools of economic policy analysis. As Michael Hanemann
(this volume) has indicated, some of these developments have been helpful while others have
not. Whether one does or does not support the market as a sole means to achieve efficiency
in water governance, it is clear that population plays a key role in achieving it as it is the
foundation for the growth of human, social and moral capital.
From an economic development perspective, it can be said that Hanemann addresses import-
ant issue regarding the values of water which are relevant for the economic growth theory and
the empirical analysis used for water management and policy. Yet, precisely in view of the
reality faced by those in developing countries as well as in the theoretical underpinnings of
the different models of economic growth, some comments have been offered regarding the
author’s assessment of the value of water and its consequences. These comments underline
the role of water use as well as the need to determine the price and cost of water, the need for
intertemporal analysis, the appropriateness of using marginal versus average values, and the
importance of taking into account institutional as well as water access conditions. All of these
issues suggest that population has a solution role in resolving some of the water management
problems that we face today.

REFERENCES

Agenor, P.R (2000). The Economics of Adjustment and Growth. Academic Press, San Diego, California,
USA.
Agenor, P.R. & Montiel, P. (1999). Economic Development. Princeton University Press, New Jersey,
USA.
Aguirre, M.S. (2002). Sustainable development: why the focus on population? International Journal of
Social Economics, 29: 12.
Barro, R. (1974). Are government bonds net wealth? Journal of Political Economy, 82(4):
1095–1117.
The value of water and theories of economic growth 101

Barro, R. (1989). Economic Growth in Cross Section Countries. Working Paper n. 3120. NBET,
Cambridge, UK.
Barro, R. & Lee, J.W. (1993). International Comparisons of Educational Attainment. Journal of
Monetary Economics, 32(3): 363–394.
Becker, G. (1974). A Theory of Social Interactions. Journal of Political Economy, 87(4): 1063–1093.
Becker, G. (1991). A Treatise on the Family. Harvard University Press, Cambridge, UK.
Becker, G. & Barro, R. (1988). A Reformulation of the Economic Theory of Fertility, Quarterly Journal
of Economics, 103(1): 1–25.
Becker, G.; Murphy, K. & Tamura, R. (1993). Human Capital, Fertility, and Economic growth. In:
G. Becker, Human Capital: a theoretical and empirical analysis, with a special reference to Education.
Chapter XII. Third Edition. Chicago University Press, Chicago, USA.
Birdsall, A. (1995). Economic Approaches to Growth and Development. In: H. Chenery & T. Srinivasan
(eds.), Handbook of Development Economics. North-Holland, Amsterdam, the Netherlands.
Blackburn, K. & Ravn, M.O. (1993). Growth, Human Capital Spillovers and International Policy
Coordination. Scandinavian Journal of Economics, 95(4): 495–515.
Brown, L.; Gardner, G. & Halweil, B. (1999). Beyond Malthus: Nineteen Dimensions to the Population
Problem. Worldwatch Institute. Washington D.C., USA.
Coale, A. & Hoover, E. (1958). Population growth and economic development in low-income countries.
Princeton University Press, New Jersey, USA.
De Long, J.B. & Summers, L.H. (1993). How strongly do developing economies benefit from equipment
investment? Journal of Monetary Economics, 32(2): 395–415.
Denison, E. (1985). Trends in America Economic Growth, 1929–1982. Brookings Institute, Washington,
USA.
Dobson, A.P.; Bradshaw, A.D. & Baker, A.J.M. (1997). Hopes for the Future: Restoration Ecology and
Conservation Biology. Science, 277 (25 July): 515–522.
Dorwick, S. & Nguyen, D.T. (1989). OECD Comparative Economic Growth, 1950–1985: catch-up and
convergence. American Economic Review, 79(5): 1010–1030.
Easterly, W.; Kremer, M.; Pritchett, L. & Summers, L.H. (1993). Good policy or good luck? Country
Growth Performance and Temporary Shocks. Journal of Monetary Economics, 32(3): 459–483.
Eberstadt, N. (1995). Tyranny of Numbers: Mismeasurement and Misrule. American Enterprise Institute,
Washington, D.C., USA.
Eberstadt, N. (2000). Prosperous Paupers and Other Population Problems. Free Press, Washington,
D.C., USA.
Ehrlich, P. (1968). The Population Bomb. Ballantine Books, New York, USA.
Ehrlich, P. & Ehrlich, A. (1990). The Population Explosion. Simon & Schuster, New York, USA.
Fukuyama, F. (1999). The Great Disruption. The Free Press, New York Fullerton, USA.
Fullerton, D. & Stavins, R. (1998). How Economists see the Environment. Nature, 395 (October):
433–434.
Furedi, F. (1997). Population and Development. St. Martin’s Press, New York, USA.
Hardin, G. (1968). The Tragedy of the Commons. Science, 162: 1243.
Hardin, G. (1998). The Ostrich Factor. Oxford University Press, Oxford, UK.
Johnson, G. (2000). Population, Food, and Knowledge. American Economic Review, 90(1): 1–14.
Kiester, E. (Jr.) (1999). A Town Buries the Axe. Smithsonian Magazine, 30(4) (July): 70–79.
King, R.G. (ed.). (1993). National Policies and Economic Growth: a World Bank Conference. Journal
of Monetary Economics. Special issue, 32(3): 359–575.
Kliksberg, B. (2000). The role of Social Cultural Capital in the Development Press. Latin American
Studies Center. College Park, University of Maryland, USA.
Kliksberg, B. (2001). The Social Situation of Latin America and its impact on Family and Education.
Organization of the American States. Washington, D.C., USA.
Komives, K.; Whittington, D. & Wu, X. (2003). Infrastructure coverage and the poor: a global perspective.
In: P.J. Brook & T.C. Irwin (eds.), Infrastructure for poor people: public policy for private provision.
The World Bank: 77–123.
Matson, P.A.; Parton, W.J.; Power, A.G. & Swift, M. (1997). Agricultural Intensification and Ecosystem
Properties. Science, 277 (25 July): 504–509.
Rabkin, J. (1997). Greenhouse Politics. American Enterprise Institute, Washington, D.C., USA.
102 M.S. Aguirre

Razin, A. & Ben-Zion, U. (1975). An intergenerational model of population growth. American Economic
Review, 65(2): 923–933.
Rosemberg, N. & Birdzell, L.E. (1986). How the West Grew Rich: The economic Transformation of the
Industrial World. Basic Books. New York, USA.
Scully, G. (1988). The Institutional Framework and Economic Development. Journal of Political
Economy, 96(3): 652–662.
Sen, A. (1981). Poverty and Famines: an Essay on Entitlement and Deprivation. Clarendon Press,
Oxford, UK.
Sen, A. (1994). Population and Reasoned Agency: Food, Fertility, and Economic Development. In:
K. Lindhal-Kiessling & H. Landberg (eds.), Population, Economic Development, and the Environment.
Sage Publications, New Delhi, India.
Simon, J. (1992). Population and Development in poor Countries: Selected Essay. Princeton University
Press, New Jersey, USA.
Simon, J. (1996a). The Ultimate Resource 2. Princeton University Press, New Jersey, USA.
Simon, J. (1996b). The State of Humanity. Blackwell, New York, USA.
Solow, R. (1956). A contribution to the theory of economic growth. Quarterly Journal of Economics,
70(1): 65–94.
UNEP (1992). Report of the United Nations Conference on Environment and Development: Agenda 21.
E.93.I.8 (Agenda 21). 3–14 June, Rio de Janeiro, Brazil.
Willis, R. (1985). A Theory of the Equilibrium Interest Rate in an Overlapping Generations Model:
Life Cycle, Institutions and Population Growth. Discussion Paper 85–8. Population Research Center,
NORC and University of Chicago, USA.
Wolfgram, A. (1999). Population, Resources, & Environment: a Survey of the Debate. Web Page: (http://
faculty.cua.edu/aguirre).
III

Irrigation
CHAPTER 6

Irrigation efficiency, a key issue: more crops per drop

K.D. Frederick
Resources for the Future, Washington, D.C., USA (currently retired)

ABSTRACT: Since the start of the green revolution in the 1960s, irrigation has accounted for about
80% of the increase in agricultural production. Irrigators now use about 70% of all water withdrawals and
80% of the water consumed worldwide to grow 40% of the world’s food and fiber. Current water uses,
however, are depleting and degrading water resources and creating doubts as to the adequacy of supplies
to meet future demands. About 10% of the world’s agricultural food output depends on non-renewable
groundwater supplies.
Existing water uses and irrigation developments emerged largely in an era when water was not viewed
or treated as a scarce resource and the environmental impacts of water projects and diversions were
ignored. Salinity and water logging have severely affected 20–30 million irrigated hectares and adversely
impacted another 60–80 million. Sedimentation is reducing storage capacity of existing reservoirs.
And the combination of high financial and environmental costs of developing new supplies and the
growing demands of non-agricultural water uses limit development of new irrigation water supplies.
A growing appreciation for the services provided by aquatic resources and concerns about irrigation’s
adverse impacts on these resources are additional obstacles to expanding agricultural as well as other
non-environmental water uses.
Past trends as well as current irrigation water uses are unsustainable. Yet, even under optimistic
projections for increasing dryland farming, continued growth of irrigated production is required to
meet future food and fiber demands. Global population is expected to increase by about 2000 million
people in the next quarter century with almost all of the increase coming in developing countries where
malnutrition is common.
The opportunities for and obstacles to reducing irrigation water use while increasing production to
avert future water and agricultural crises are examined. Potential opportunities include institutional
reforms to provide incentives for more efficient water use and development of higher yielding and more
nutritious plant varieties, crops that can be grown in more hostile environments or with lower quality
water, and varieties that yield more harvestable biomass per unit of water. Potential obstacles include
vested interests that benefit from current water laws and inefficiencies, concerns about the potential
environmental impacts of genetically modified food, and the high costs of the research and infrastruc-
ture investments needed to make these potential opportunities possible. The implications of increasing
levels of atmospheric CO2 on plant photosynthesis and future climate conditions add uncertainty to the
challenge of meeting the agricultural demands of a larger and, hopefully, more affluent and better-fed
population in an environmentally benign and sustainable way.

Keywords: Agricultural water use, irrigation efficiency, water scarcity and agricultural production,
climate change, irrigation water

1 THE ROLE OF IRRIGATION

Irrigation is the dominant use of water and essential to the production of the world’s food and
fiber. Irrigators use about 70% of all water withdrawn from streams, lakes, and groundwater
106 K.D. Frederick

aquifers and about 80% of the water consumed worldwide. The 250 million irrigated hectares
produce 40% of the world’s food and fiber on only 17% of all cropland (Shortle & Griffin,
2001: ix).
Irrigation in combination with high yielding crop varieties and the application of fertilizer
and pesticides have accounted for about 80% of the increase in agricultural production since
the start of the green revolution in the 1960s. Synergies between irrigation and the green
revolution technologies contributed to this growth. Reliable supplies of water are essential for
achieving most of the benefits of these technologies. And the higher yields that are possible
with these technologies encourage the spread of irrigation.
The 5-fold increase in irrigated land in the last century has stressed water supplies in
many areas around the world, creating doubts about the adequacy of supplies to meet future
agricultural as well as other demands for water. The combination of its current water use, its
importance to meeting current and future demands for food and fiber, and concerns about the
sustainability of agricultural water uses give irrigation a special role in determining whether
a water crisis is or will become a myth or reality.

2 SIGNS OF STRESS

Current water uses are depleting and degrading some water resources. Water diversions
for irrigation and other uses have severely degraded many rivers and lakes around the
world. Salinity and water logging from poor drainage of irrigated lands pose additional
problems for meeting projected water and agricultural demands. The Food and Agriculture
Organization (FAO) estimates that salinity and water logging have severely affected 20–30
million irrigated hectares and adversely impacted another 60–80 million hectares (Rosegrant
et al., 2002a: 2).
About 10% of the world’s agricultural food output now depends on non-renewable ground-
water supplies. Water tables are falling a meter or more annually in parts of Mexico, India,
China, the USA, and several other countries (World Commission on Water for the 21st Century,
2000: 287). In the USA, groundwater use for irrigation exceeds recharge on about 4 million
hectares, 20% of its irrigated land. Rising water costs due to declining well yields and increas-
ing pumping depths have reduced irrigation from the Ogallala aquifer in the USA High Plains
by more than a million hectares since the mid-1970s. Projections indicate that rising water
costs will force more than 2 million irrigated hectares in this region to be abandoned or returned
to dryland farming by the year 2020 (National Research Council, 1996: 130–133).
The ecological, economic, and human health damages attributable to the expansion of
irrigation in the Aral Sea basin are extreme but not isolated examples of the costs of excessive
water diversions, poor drainage, and runoff contaminated by agricultural chemicals. Since
1960 the sea has lost about two-thirds of its volume and become too saline to support the
former thriving fishing industry. The productivity of the delta ecosystems of the basin’s two
main rivers has suffered. High salt levels forced about 1 million hectares out of production
and reduced yields an estimated 60% on the remaining lands. And contamination of drinking
water supplies by pesticides and toxic salts blown by dust storms from the exposed seabed have
increased mortality and morbidity rates in the lower basin (Frederick, 1991; Postel, 1999).
The costs of developing additional water supplies are rising for several reasons. Dams and
reservoirs have been the primary means of converting naturally varying water resources into
Irrigation efficiency, a key issue: more crops per drop 107

more reliable and controlled supplies. However, there are limits to and diminishing returns to
the safe yield produced by successive increases in a river’s reservoir capacity. Because the best
reservoir sites are developed first, subsequent increases in storage require larger investments.
For example, a study of decadal changes in reservoir capacity produced per unit volume of
dam for the 100 largest dams in the USA suggests that average capacity per m3 of dam declined
35-fold from the 1920s to the 1960s (United States Geological Survey, 1984: 33). A stream’s
maximum possible yield is limited by its average annual flow. But reservoir evaporation losses
offset the gains from surface storage well before this maximum is reached. The social costs of
storing and diverting water also increase as the number of free-flowing streams declines and
society attaches more value to water left in a stream.
Sedimentation is estimated to be reducing reservoir storage capacity by about 1% per year.
Replacing this capacity with new reservoirs might cost US$ 10,000–13,000 million a year
if enough reservoir sites could be found. But dredging sediment to preserve the capacity of
existing reservoirs could be 10-times more expensive according to Postel (1999: 86). In the
USA, reservoir losses from sedimentation have probably exceeded additions to storage from
new construction since 1990 (Frederick & Gleick, 1999: 28).
Evidence of diminishing returns and rising water costs in the developing countries is abun-
dant. Irrigation costs more than doubled from 1970 to 1990 in India, Indonesia, and Pakistan.
Costs increased 3-fold in Sri Lanka, more than 50% in the Philippines, and 40% in Thailand
in recent decades. High costs combined with declining cereal prices result in low economic
returns for new irrigation projects in Asia. Developing more water for irrigation is likely to be
even less attractive economically in Latin America and Africa where construction costs tend
to be higher than those in Asia (Rosegrant et al., 2002a: 3–4).
Wetlands are valuable components of hydrologic systems. They purify water by filtering
and settling pollutants and sediments, and they reduce flooding by dispersing high water flows
over time and area. They are among the earth’s most productive ecosystems, providing habitat
and food for fish and wildlife and a variety of harvestable resources such as timber, berries,
fish, fur, and peat. Irrigation projects, however, have contributed to the loss of hundreds of
millions of hectares of wetlands that have been drained and filled for cropland, flooded behind
dams, and dried up as a result of streamflow depletion.
Existing water uses and irrigation developments emerged largely in an era when water was
not viewed or treated as a scarce resource and the environmental impacts of water projects and
diversions were ignored. But a growing appreciation for the services and amenities provided by
wetlands, streams, lakes, and other aquatic resources and concerns about irrigation’s adverse
impacts on these resources are now additional obstacles to expanding agricultural as well as
other offstream water uses. The combination of high financial and environmental costs of
developing new supplies and the growing demands of non-agricultural water uses constrain
development of new irrigation water in many areas of the world.
Pressures to reallocate supplies from irrigation, which is a relatively low-value water use,
to other uses are increasing as the demand for non-agricultural water rises with population
and income growth and as water becomes scarcer and more costly. Globally, withdrawals for
domestic and industrial uses rose 4-fold from 1950 to 1995, nearly twice as fast as agricultural
uses (Rosegrant et al., 2002a: 2). To date, transfers of water supplies once used for agri-
culture to domestic, industrial, or environmental uses have posed little problem to meeting
growing food and fiber demands. Adoption of more efficient irrigation practices and green
revolution technologies have produced more crops per drop. To meet future agricultural and
108 K.D. Frederick

water demands, continued increases in the productivity per unit of water for both dryland and
irrigated agriculture will be essential.

3 THE CHALLENGE

Global population, about 6000 million at the start of this century, is expected to reach nearly
8000 million by the year 2025. Almost all of this growth is expected to occur in the developing
countries where 1300 million people currently lack access to safe water, 2600 million lack
sanitation facilities, malnutrition is common, and average per capita caloric consumption is
less than 80% of the average for the OECD countries. Increasing agricultural production to
meet the demands of a larger and, hopefully, more affluent and better-fed population in an
environmentally benign and sustainable way is an imposing task in view of the water and
agricultural stresses detailed above.
The World Commission on Water for the 21st Century (2000) concluded that we are on
an unsustainable path leading toward a water crisis. The linkages between water and agri-
culture suggest this could also be a path toward a food and health crisis for much of the
developing world. With business as usual being unsustainable and likely to result in unaccept-
able production and environmental outcomes, the challenge is to identify an alternative path
that is sustainable and will meet growing water and agricultural demands and then to adopt
the policies required to move to this path. The Impact-Water model developed by Rosegrant
et al. (2002a) described below attempts to identify a strategy that will achieve environmental
sustainability and meet the water and food demands of the larger and more affluent population.

4 THE IMPACT–WATER MODEL

Rosegrant, Cai, and Cline developed a global model of water and food supply and demand to
examine the implications of alternative water and investment policies on water supplies and
use, agricultural production, and the environment between 1995 and 2025. Global demand
for cereals and meat is projected to grow by 46% and 56% respectively over these 30 years.
In the developing countries, which are expected to account for more than 80% of the world’s
population by 2025, demand is projected to increase by 65% for cereals and more than double
for meat (Rosegrant et al., 2002a: 2).
The Impact-Water model (IM) combines an international model of agricultural produc-
tion and trade with a water simulation model. The agricultural Impact model examines the
effects of various food policies and rates of agricultural research on crop productivity, and the
impacts of income and population growth on food supply and demand. The supply, demand,
and prices for 16 agricultural commodities are determined in 36 regional or country submod-
els that are linked through trade. The impacts of water availability on agricultural production
are introduced through the water simulation model that accounts for renewable water sup-
plies, non-agricultural water demand, water infrastructure, and economic and environmental
policies related to water development and management. Water availability is introduced as
a stochastic variable with observable probability distributions based on the 30-year climate
record from 1961–1990. The IM model is used to consider the future of water and food in
the year 2025 with a projected world population of 7900 million, an increase of nearly 2000
Irrigation efficiency, a key issue: more crops per drop 109

million from the 1995 base year, under several alternative scenarios (Rosegrant et al., 2002a:
18–31).

4.1 Business as usual


The business as usual scenario assumes a continuation of current trends in water and food policy
and management. This scenario and its implications are described in Rosegrant et al. (2002a:
35–38, 61–108). Institutional and water management reforms are limited and piecemeal, and
government investments in agriculture and irrigation are reduced. Under this scenario, water
withdrawals are projected to increase 22% globally and 27% in the developing countries from
1995 to 2025. Total consumptive use increases 16% globally and 18% in the developing world.
Irrigation consumptive use rises only 4% and accounts for only 20% of the global increase.
Water scarcity intensifies in the developing world, which accounts for 93% of the total
increase in water consumed for irrigation. Economic incentives encourage some farmers
to adopt more efficient irrigation practices. But political opposition from those concerned
about the impacts of higher water prices and entrenched interests that benefit from existing
water allocation deters adaptation in some areas. Overall, irrigation efficiency and river basin
management improve slowly. Groundwater mining continues and there is no change in the
share of water for environmental uses.
Water scarcity leads to slower growth of food production and shifts in where it is grown.
Irrigated and rainfed agriculture each account for about half of the production increase. Greater
food production depends largely on higher yields because of the slow growth in the area devoted
to food crops. But growth in yields declines from the 1.5% average from 1982 to 1995 to 1%
from 1995 to 2025 because many of the actions that produced past yield grains are not easily
repeated and public investment is reduced. The growth of food production in developing
countries slows as a smaller fraction of their irrigation water demand is met. Relative crop
yield (the ratio of actual projected yield to the economically attainable yield without water
stress) for cereals in the developing countries declines from 0.86 in 1995 to 0.75 in 2025. The
decline attributable to increased water stress is equivalent to an annual production loss of 130
million metric tons, China’s annual rice production in the late 1990s. Consequently, developing
countries become much more dependent on food imports under the business as usual scenario.

4.2 A crisis scenario


A moderate worsening of current trends in water and food policy and investment results in
a genuine water crisis according to the authors of the IM model. The water crisis scenario
and its implications are described in Rosegrant et al. (2002a: 38–40, 109–136). Deteriorating
infrastructure and poor management lead to a decline in water-use efficiency. Productivity
growth in rainfed areas declines, particularly in marginal areas. Erosion and sediment loads
in rivers rise as upper watersheds are deforested by people forced to turn to slash-and-burn
agriculture. More reservoir storage is lost to sedimentation, wetland losses increase, and the
health of aquatic ecosystems is further compromised as the water reserved for environmental
purposes declines. Groundwater mining accelerates until about 2010 when declining well
yields and rising pumping depths start to make irrigation from some key aquifers in China,
India, and North Africa too expensive. In comparison to the business as usual scenario, global
water consumption is 13% higher in 2025, total cereal production is 10% lower, food prices are
sharply higher, and food security in the developing countries deteriorates. Irrigation accounts
110 K.D. Frederick

for virtually all of the increase in water use between the crisis and business as usual scenarios
as farmers attempt to compensate for lower efficiency and water losses.

4.3 Toward sustainability


A sustainable water outcome under the IM model requires increased investment in crop
research, technological change, and water management reform. The sustainable water use
scenario and its implications are described in Rosegrant et al. (2002a: 40–44, 109–136). Water
prices to the agricultural sector are gradually increased, providing incentives to conserve and
funds to maintain and build new infrastructure. Water markets provide irrigators opportunities
to profit from conserved water and reallocate supplies to alternative uses. Communities and
water user associations are given responsibility for operating and managing irrigation sys-
tems. On-farm investments in irrigation and water management technology increase along
with the efficiency of both irrigation systems and basin wide water use. Higher water prices
for municipal and industrial uses also encourage investments in conservation and recycling.
Groundwater overdrafts are gradually phased out through regulations, stricter enforcement,
and market-based approaches that assign groundwater rights.
In comparison with business as usual, global water consumption is 20% lower by the year
2025 under the sustainable water scenario. Irrigation use also declines by 20% and accounts
for nearly three-fourths of the overall reduction. The increased priority given to environmental
water uses initially reduces the reliability of water for irrigation. Over time, however, more
efficient water use helps offset the loss of water and the reliability of irrigation water supplies
improves. But irrigated cereal production under the sustainable scenario is 5% lower than with
business as usual.
Total cereal production in 2025 under the sustainable scenario is 48% above the 1995
baseline and 1% higher than business as usual. Irrigated cereal production rises 49% over
the 30 years in spite of a 17% decline in the amount of water consumed for irrigation. The
developing countries increase irrigated cereal production 50% consuming 19% less water. The
developed countries produce 46% more cereals using 5% less water.
Underlying the 1% difference in cereal production between the sustainable and business
as usual scenarios are significant changes in how and where crops are grown. Global cereal
yields are 7% higher on rainfed lands and 2% lower with irrigation due to the lower reliability
of its water supply in comparison to the business as usual scenario. Cereal production is 29
million tons higher in the developing countries and 10 million tons lower in the developed
countries, implying a reduction in net food trade under the sustainable scenario.
Rosegrant et al. (2002b: 16) conclude that the IM model sustainability scenario “shows
that with improved water policies, investments, and rainfed cereal crop management and
technology, growth in food production can be maintained while universal access to piped
water is achieved and environmental flows are increased dramatically”. Improved productivity
of rainfed agriculture is critical to meeting agricultural demands and reducing the pressures
on water supplies. Under the sustainable scenario, rainfed agriculture accounts for 57% of
the projected increase in global cereal production from 1995 to 2025. This is in contrast to
developments over the last third of the 20th century when the spread of irrigation and the green
revolution technologies accounted for about 80% of the increase in agricultural production.
Infrastructure investments that link remote farmers to markets and policy changes that reduce
the risks of rainfed agriculture and encourage on-farm investments contribute to dramatic
yield increases even in drought-prone and high-temperature areas.
Irrigation efficiency, a key issue: more crops per drop 111

Research expands the opportunities for increasing crop and water yields on both irrigated
and dryland farms and institutional reforms provide the incentives to conserve water and adopt
better agricultural and water management practices. Advances in water harvesting systems
and adoption of advanced farming techniques such as contour plowing, land leveling, and
minimum-till and no-till technologies enable more effective use of rainfall in crop production.
These improved farming techniques result in less erosion and reservoir sedimentation, reducing
the need for new reservoir capacity.
The major role of dryland production projected in the sustainable water use scenario reduces
the changes in irrigation required to ensure food security and water sustainability. But even
with the increased reliance on dryland farming, global irrigated cereal production is projected
to grow by nearly 50% over the 30-year period while consuming 17% less water (Rosegrant
et al., 2002a: 110–123).

5 SUSTAINABILITY: OPPORTUNITIES AND OBSTACLES

With irrigation consuming about 80% of all water withdrawn worldwide and current uses
depleting and degrading supplies, adoption of more water-conserving and environmentally-
benign irrigation practices are critical for achieving sustainable water use. On the other hand,
meeting future food and fiber demands requires continued growth of production from irrigated
lands even under the IM model’s optimistic projections for increasing dryland agricultural
output. Postel (1999: 10) concludes that future production increases will depend more on
irrigated than on dryland farming. She is skeptical about the prospects for achieving large
increases from dryland farming because most of the agricultural lands endowed with abundant
and reliable rainfall, such as the USA cornbelt and Western Europe’s wheat areas, are already
producing close to their maximum potential.
The opportunities for and obstacles to reducing irrigation water use while increasing
production sufficiently to avert future water and agricultural crises are examined below.

5.1 Water use efficiency


Irrigation involves withdrawing water from a surface or groundwater source and delivering it to
a plant. The portion consumed in the irrigation process consists of water that is: (1) transpired
by the plant; or (2) evaporated from soil and plant surfaces and lost as a result of infiltration
and runoff to locations where it cannot be economically recovered. The portion of the water
withdrawn that is not consumed adds to groundwater recharge or downstream supplies and is
available for other uses.

5.1.1 Plant transpiration


Transpiration is affected by many factors, including plant type and cover, stomatal behavior,
and atmospheric carbon dioxide concentrations. The irrigation method employed to deliver
water to the crops usually does not affect transpiration unless it fails to provide the plant
with sufficient supply on a timely basis. Reducing plant transpiration usually results in
water stress and reductions in the quantity and perhaps the quality of the yield (National
Research Council, 1996: 106). Researchers have consistently found a linear relationship
between a crop’s transpiration and its yield of plant matter up to the point at which water
is no longer a limiting factor. This suggests that for any given crop in a given location,
112 K.D. Frederick

more plant production results in more water consumed through transpiration (Postel, 1999:
172–173).
There are differences in the efficiency with which plants use water. Consequently, changes
in the crops that are grown and development of new varieties might produce more crop per
drop. Moreover, exposing plants to higher concentrations of carbon dioxide increases their
rate of photosynthesis, which normally leads to higher yields, and can reduce the amount of
water transpired by the plant (Mendelsohn & Rosenberg, 1994: 24).
Plants such as corn, sorghum, millet, and sugar cane (C4 plants) are naturally more efficient
photosynthesizers than C3 plants, which include most of the world’s small grains, legumes,
root crops, cool season grasses, and trees. Laboratory and greenhouse studies indicate that
doubling the level of carbon dioxide increases the yields of C3 plants by 15% to 20% and
of C4 plants by 5% (Adams et al., 1999: 12). The enriched carbon environment also results
in partial closure of a plant’s stomates (pores), reducing transpiration from a given leaf area.
If these results hold up under open field agriculture, significant gains in output per unit of
water might result from the increases in atmospheric carbon dioxide concentrations that are
anticipated over this century.
A number of complicating factors could offset these potential gains in water efficiency.
The interaction between temperature levels and evapotranspiration is particularly relevant
since carbon dioxide is the principal greenhouse gas contributing to global warming. Shugart
et al. (2003: 18) note that “even if increased atmospheric CO2 could potentially allow a tree
to keep the stomata of its leaves closed for longer periods, it might still need to continue to
leave the stomata open for the purpose of evaporative cooling to maintain heat balance. As a
consequence, the improvements in water use efficiency could be offset by the need to maintain
heat balance”.

5.1.2 Irrigation efficiency


Irrigation water losses to evaporation, infiltration, and runoff are common and often large
at the farm level. In developing countries on-farm irrigation efficiency is generally between
20–40%. However, water-use efficiency for the basin as a whole can be much higher than
on individual farms because some of the water that is not consumed by evapotranspira-
tion at the farm can be an important source of supply for downstream users. In Egypt’s
Nile basin, for example, overall basin efficiency is estimated at 80% even though indi-
vidual water efficiency averages only 30% (Rosegrant et al., 2002a: 6). Because of these
and other externalities associated with irrigation practices, the benefits and costs of alter-
native irrigation systems and cultivation practices are likely to differ at the farm and basin
levels.
On-farm water losses and perhaps higher crop yields can be achieved through investments
in more efficient irrigation systems and adoption of improved management practices. Lining
canals or installing gated pipe can reduce losses in delivering water from the farm gate to the
head of the field. Surface irrigation systems such as flood and furrow are the most common
methods of delivering water to crops. With these systems large quantities of water are usually
applied to ensure the entire field is irrigated. More than half of the water may run off the field
or infiltrate into the soil beyond the root zone of the plant. Land leveling, shorter furrows, and
construction of borders enable a more uniform distribution of water to the crops, reducing
on-farm water losses and the amount of water required to ensure all the plants are irrigated.
Sprinkler systems can reduce or eliminate surface runoff, but they may increase evaporation
Irrigation efficiency, a key issue: more crops per drop 113

losses in a hot windy environment. Drip irrigation applies water directly to the root zone of
the crop, which can virtually eliminate field water losses.
Irrigation systems that apply water more uniformly to the field can increase crop yields as
well as reduce on-farm water use. For example, studies in India, Israel, Jordan, Spain, and
the USA show that drip irrigation has increased crop yields by 20–90% with 30–70% less
water (Postel, 1999: 174). The state of the art for increasing the amount of crop per drop
would combine microirrigation with sensors to monitor soil moisture and computer programs
that analyze when and how much to irrigate. Since drip and other highly efficient irrigation
systems are used on only 1% of the world’s irrigated area, the possibilities for increasing
yields to water are large. However, the economic, management, and institutional obstacles to
widespread adoption of these methods are also large.
Increasing on-farm irrigation efficiency usually implies higher costs as well as benefits to
the farmer. The net benefits or costs to the farmer of investing in a given irrigation system
or tillage practice vary depending on factors such as soil conditions, capital and labor costs,
management skills, and water costs. The net social and private benefits of the investment
will differ depending in part on its impacts on the quantity and quality of water available to
downstream users. Investments that reduce consumptive use on the farm while maintaining or
increasing agricultural output increase effective water supplies. Moreover, any practice that
increases irrigation efficiency generally decreases the sediment, salts, and chemicals that can
pollute downstream supplies. Reducing erosion helps protect a farm’s long-term productivity
as long as salts do not accumulate in the root zone.
In the short run, drip irrigation might be able to reduce on-farm consumptive water use to
little more than the amount transpired by the plant. However, sustainable irrigation requires
applying enough additional water to leach salts from the root zone and drainage to remove
water from beneath the field. All surface and groundwater supplies contain salts that are
left behind when water is transpired by plants and evaporated from fields. Allowing salts to
accumulate in the soil is detrimental to plant growth. Inadequate drainage results in rising
groundwater tables and eventually to waterlogging, which destroys the field’s productivity by
cutting off a source of oxygen to the plants and by concentrating salts near the surface.

5.2 Biotechnology
The hybrid seed varieties that led to the green revolution and the rapid expansion of irri-
gated agriculture starting in the 1960s increased the output per unit of land and water largely
by increasing the harvestable proportion of a plant’s biomass. According to Postel (1999:
192–193), many plant breeders believe that the most fruitful possibilities for increasing the
harvestable proportion of the major irrigated grains have already been exploited.
Advances in genetic engineering in recent decades have generated hopes that genetic
modification of plants will be a panacea for many of the world’s agricultural problems and
provide the tools for another green revolution. Technological optimists anticipated more than
a decade ago that biotechnology would lead to the rapid development and introduction of
new higher-yielding and more nutritious plant varieties, crops that can be grown in more
hostile environments or with lower quality water, and varieties that yield more harvestable
biomass per unit of water. Economic considerations and concerns about the potential environ-
mental impacts of genetically modified (GM) crops have dampened some of the initial high
expectations for these crops.
114 K.D. Frederick

The area planted with GM crops has grown from virtually nothing in the mid-1990s to
about 60 million hectares, roughly 4% of the world’s arable land. To date, however, GM crops
have had little, if any, impact on how much crop is produced per drop or the plant’s ability
to utilize lower quality water supplies. To justify the costs of developing and testing new
GM plants, companies have focused their research on developing seeds farmers will purchase
new every year. Four crops – maize, soya, canola, and cotton – account for virtually all the
GM planted area. And the major contribution of the GM seeds from a farmer’s perspective
is that they provide greater protection against insects and herbicides, reducing the need to
purchase agricultural chemicals. When farmers in rich countries are being paid to take land
out of cultivation and irrigation water around the world is subsidized, there is little incentive
to develop genetically altered crops that provide resistance to drought or that can be grown
using brackish and saline water (The Economist, 2003).
Economic factors are not the only obstacle to the development and introduction of GM crops
that produce more crop per drop. Since 1998 the European Union has banned the planting
or importation of any new genetically engineered crop. Environmental activists have trashed
fields planted with GM crops. Despite widespread hunger, Zambia has refused to accept
USA food that contained genetically modified corn. And Mexico has prevented CIMMYT,
an international organization dedicated to creating better crops for farmers in the developing
world, from testing GM corn outside a greenhouse (Charles, 2003).
Although current conditions may not encourage investments designed to develop seeds
that increase the crop per drop, this should change as the need to produce more food and
fiber with less irrigation water mounts. In spite of the lack of strong economic incentives,
researchers have identified genetic traits involved in protecting plants from salt, cold, and
drought. As water becomes scarcer and more expensive and GM crops more widely accepted,
these characteristics are likely to be incorporated into plants suitable for field application. The
holy grail for meeting future agricultural demands with less water would be to bioengineer
plants to photosynthesize water more efficiently without sacrificing other desirable traits of the
plant. Research on one of the proteins involved in photosynthesis has improved its productivity
in lab tests. Some scientists, however, are pessimistic about our ability to eventually alter
photosynthesis, a process that has remained virtually unchanged for 2000 million years of
plant evolution and natural selection (Postel, 1999: 192–193; The Economist, 2003).

5.3 Institutional reforms


The institutions that establish the opportunities and incentives to use and abuse water resources
are largely relics of an era when water was viewed as a free resource. They currently inhibit
the development and adoption of water conserving technologies.
Irrigators are often the principal beneficiaries of the laws and subsidies influencing water
use. They are the largest users and have high-priority rights for much of the readily accessible
supplies. They pay nothing for the water itself, and the costs of the projects that store and deliver
water to their farms are often subsidized. Farmers receiving water from government projects
rarely pay more than 20% of water’s real cost and generally much less (Postel, 1999: 230).
Low prices and lack of opportunities to profit from conserved water provide little incentive to
invest in efficient irrigation systems.
There is a growing consensus that greater reliance on economic principles in managing
and allocating water is critical for more efficient and sustainable use. The report of the World
Irrigation efficiency, a key issue: more crops per drop 115

Commission on Water for the 21st Century (2000) advocates adopting full cost pricing, the
polluter and user pays principle, and water marketing to encourage water conservation research
and investments. Higher water prices are a critical element of the sustainable water scenario
developed by Rosegrant, Cai, and Cline described above. Under this scenario, agricultural
water prices would be twice as high in developed countries and three times as high in developing
countries by 2025 compared to business as usual.
Raising water prices and reducing subsidies can be politically risky. Water is commonly
viewed as too important to be subjected to market forces, and attempts to eliminate or reduce
agricultural subsidies have generally failed in the face of strong resistance. A politically more
acceptable approach for introducing economic incentives to conserve water is to provide
irrigators with opportunities to sell unused supplies. They would then value water in terms of
its opportunity cost – the value they could get by selling water – rather than at the subsidized
price they pay for it. Selling saved water would provide funds for water-conserving and yield-
increasing investments.
Charging the full costs of water and developing active water markets also requires over-
coming practical barriers such as lack of equipment to measure how much is delivered or to
terminate deliveries to non-paying farmers. Since it is expensive to move water out of exist-
ing conveyance facilities, the current infrastructure limits the opportunities to economically
transfer water. The laws and regulations that define water rights and limit where and how it is
used may also need to be changed to establish viable water markets.
While water markets can be an important step for introducing economic principles into irri-
gation decisions, they are not a panacea for achieving efficient use. Efficient markets require
well-defined, transferable property rights, and the full costs and benefits of a transfer must be
borne by the buyer and seller. Both the nature of the resource and the institutions that allocate
and manage water can make it difficult to meet these conditions. Transferring water from one
use or location to another often affects third parties as well as the buyer and seller. The variety
of public as well as private services it provides and the interdependencies among users limit the
potential for establishing unfettered and efficient water markets. Nevertheless, in view of the
current inefficiencies and lack of incentives to conserve, most any changes that facilitate trans-
fers in response to changing supply and demand conditions and provide irrigators with incen-
tives to conserve are likely to result in a more socially efficient water use (Frederick, 2001).

5.4 Global warming


The prospect of a greenhouse warming is another reason for introducing institutional changes
that facilitate voluntary water transfers. Global warming would alter precipitation, evapotran-
spiration, and runoff patterns, as well as the magnitude and frequency of extreme events.
However, the nature, timing, and even the direction of the impacts on the regional and local
scales of primary interest to water planners and managers are uncertain. Some regions are
likely to experience more intense precipitation days, as well as increased flooding. Others
may have more frequent and severe droughts. And some may experience both more intense
floods and droughts.
The impacts of a greenhouse warming on water supplies are likely to create winners as well
as losers. On balance, however, the impacts are likely to be negative in part because water-
supply systems are designed and operated on the assumption that future climate will look like
the past climate. The prospect of an unusually rapid change in the climate adds uncertainty
116 K.D. Frederick

and risk to the design and operation of new water projects. Consequently, developing institu-
tions that facilitate and encourage voluntary water exchanges through markets would provide
a system that is both more efficient and better able to adapt to whatever the future might
bring.

6 CONCLUSIONS

A recent report of the Food and Agriculture Organization (FAO) concludes that chronic hunger
increased in the last half of the 1990s to a total of 842 million people. About 95% of these
people live in poor, developing countries (Lynch, 2003). Poverty, disease, wars, and agricul-
tural trade policies, not global food production, are the major causes for the recent increases
and current levels of chronic hunger in the world. Indeed, overproduction and lack of demand
have characterized agricultural markets in the USA and Europe in recent decades. Neverthe-
less, meeting the food and fiber demands of nearly 8000 million, hopefully more affluent,
people in the year 2025 in an environmentally benign and sustainable way is an imposing task.
Current water use is unsustainable and likely to lead to future water and agricultural crises.
However, meeting future demands while reducing irrigation water use to sustainable levels
appears to be achievable with major changes in business as usual. Large increases in invest-
ments in research, infrastructure, and efficient irrigation systems are needed. Institutional
changes that encourage treating water as an economic good are also essential. But even with
changes that produce more crop per irrigated drop and utilize lower quality water for which
there is little demand from non-agricultural users, irrigation cannot be expected to continue
producing the vast majority of the increases in agricultural output required to meet the demands
of a growing and more affluent global population. Rainfed agriculture may need to account
for half or more of the increases in production required to balance future agricultural supplies
and demands.
Meeting future food and fiber demands depends on the quantity and quality of production
that actually reaches the consumers and producers that use the farm output. Inefficiencies
in agricultural marketing can result in large losses. Reducing transportation and storage
losses is another means of increasing the effective marketable output per unit of water
transpired.
Water and food crises tend to be regional and local in nature, rather than global. Changes in
precipitation, and temperatures associated with a greenhouse warming could exasperate future
regional problems. The ability to respond to these regional and local problems depends on
both the adequacy of global agricultural production and international trade. In Africa, which is
already burdened by political instability, extreme poverty, and poor infrastructure, agricultural
production is particularly vulnerable to climate changes. Meeting their future food needs is
likely to depend on the largess of the international community, which in the past has often
fallen short of Africa’s needs. China’s food production is also highly dependent on benign
precipitation conditions as well as non-sustainable groundwater irrigation. But in contrast to
the situation in Africa, China’s dynamic economy enables them to import any food shortfalls
as long as there is adequate global production.
Longer-term sustainable food and fiber production requires dealing with population growth
well beyond the year 2025 used for the Impact-Water model projections and adapting to the
impacts of a greenhouse warming which are ignored in the modeling results.
Irrigation efficiency, a key issue: more crops per drop 117

REFERENCES

Adams, R.M.; Hurd, B.H. & Reilly, J. (1999). Agriculture and Global Climate Change: A Review of
Impacts to U.S. Agricultural Resources. Pew Center on Global Climate Change. Arlington, Virginia.
USA.
Charles, D. (2003). Corn that Clones Itself. Technology Review, 106(2): 32–41.
Frederick, K.D. (1991). The Disappearing Aral Sea. Resources, 102: 11–14.
Frederick, K.D. (2001). Water Marketing: Obstacles and Opportunities. Forum for Applied Research and
Public Policy, 16(1): 54–62.
Frederick, K.D. & Gleick, P.H. (1999). Water & Global Climate Change: Potential Impacts on U.S. Water
Resources. Pew Center on Global Climate Change. Arlington, Virginia, USA.
Lynch, C. (2003). Chronic Hunger is Increasing. Washington Post, November 27, A20.
Mendelsohn, R. & Rosenberg, N.J. (1994). Framework for Integrated Assessments of Global Warming
Impacts. Climatic Change, 28: 15–44.
National Research Council (1996). A New Era for Irrigation. National Academy Press. Washington D.C.,
USA.
Postel, S. (1999). Pillar of Sand: Can the Irrigation Miracle Last? WW Norton & Company. New York,
USA.
Rosegrant, M.W.; Cai, X. & Cline, S.A. (2002a). World Water and Food to 2025: Dealing with Scarcity.
International Food Policy Research Institute. Washington D.C., USA.
Rosegrant, M.W.; Cai, X. & Cline, S.A. (2002b). Global Water Outlook to 2025: Averting an Impending
Crisis. International Food Policy Research Institute. Washington D.C., USA.
Shortle, J.S. & Griffin, R. (eds.) (2001). Irrigated Agriculture and the Environment. Edward Elgar
Publishing. Northhampton, Massachussets, USA & Cheltenham, UK.
Shugart, H.; Sedjo, R. & Sohngen, B. (2003). Forests and Global Climate Change: Potential Impacts
on U.S. Forest Resources. Pew Center on Global Climate Change. Arlington, Virginia, USA.
The Economist (2003). Climbing the helical staircase: A survey of biotechnology. March 29: 1–24.
United States Geological Survey (1984). National Water Summary 1983 – Hydrologic Events and Issues.
Water Supply Paper 2250. U.S. Government Printing Office. Washington D.C., USA.
World Commission on Water for the 21st Century (2000). A Report of the World Commission on Water
for the 21st Century. Water International, 25(2): 284–302.
CHAPTER 7

Irrigation efficiency, a key issue: more crops per drop.


Observations and comments

P. Arrojo
Department of Economic Analysis, University of Zaragoza, Spain

ABSTRACT: Some samples of the current unsustainability are: the breakdown in the biodiversity and
specially in fisheries; the general degradation of rivers, lakes, wetlands and aquifers; the grave social
impacts in poor communities by large dams and by the general degradation of aquatic ecosystems; impacts
on deltas, coastal areas and coastal platforms; degradation of irrigated lands diminishing productivity;
or the breakdown of the water cycle and its influence on climate changes. The growth of irrigation is
the key of the crisis of sustainability in the aquatic ecosystems, and at the same time the key in the fight
against hunger. Sustainability of ecosystems cannot be anymore considered as a restriction but a need
for fighting successfully against hunger.
There are different functions of water: water for life, water as public service and general interest,
water for business. These different functions of water ask for different management approaches. We
must recognise priorities among these functions.
This chapter also focuses on in tendencies, paradoxes and socio-economic contradictions that currently
exist in water management, related to hydraulic works, markets, subsidies, prices, or WTO (World Trade
Organization) policies.
Beyond recognising the value of some traditional production models, we need to design and promote
a new water culture (the culture of sustainability): better life with less resource (not only by technical
and market efficiency) and less and different way of consumption. How to produce can be even more
important than how much. It is necessary a new cultural approach of consumption: quality instead
of quantity. Assessment of eco-social efficiency given by integrated traditional models of living and
producing in sustainable ways is needed.

Keywords: Sustainability, eco-social efficiency, values of water, new water culture

1 THE CRISIS OF WATER ECO-SYSTEM SUSTAINABILITY

The spectacular growth in irrigated agriculture – its surface area increasing fivefold over the
course of the 20th century – has undoubtedly been one of the key factors in the fight against
hunger around the world. However, at the same time it has also been – and still is – the key
factor in the crisis of unsustainability which is putting the health of most continental water
eco-systems at risk, degrading the biological productivity of rivers, lakes, wetlands and coastal
platforms, and aggravating the food crisis it aims to resolve, thus completing the vicious circle
of unsustainability.
As Professor Frederick (this volume) indicates, 10% of food production is based on ground
waters, the aquifers of which are at their recharge limits. This is only the tip of the iceberg of
the unsustainable logic upon which we have based our present development model.
120 P. Arrojo

In his chapter, Professor Frederick deals with some of the clearest and best known cases
of unsustainability, such as the case of the Aral Sea, or the large aquifers of Mexico, India,
China or the USA. However, as an indicator of the dynamics of this unsustainable exploitation,
we should also note the breakdown in the bio-diversity of continental aquatic ecosystems and
the serious effects on marine fisheries, owing to the breakdown of these continental water
systems.
The very concept of sustainability for the continental aquatic ecosystems is currently under
debate and still being agreed upon, from both a biological (sciences of nature) and also a social
perspective (corresponding commitments). However, it seems evident that from this dynamic
process, the trend points towards diagnosing the current degradation of our rivers, lakes,
wetlands and aquifers as being unacceptable, chiefly owing to the dumping of pollutants and
excessive water extraction. Beyond ecological degradation, if we consider the serious effects
on human health and on the basic resources of many communities, as well as the undermining
of the human rights of entire settlements affected by large dams (between 40 and 80 million
people have been forcefully evicted from their homes, according to estimates in the World
Commission on Dams report (WCD, 2000), we can see a clearer picture of the eco-social
complexity of unsustainability.
The most dramatic ecological, social, and environmental problems for continental aquatic
ecosystems occur in arid or semi-arid steppe areas, as a result of excessive diversions. However,
it is also true that in more humid climates, impacts on the continental hydrological cycle are
produced by a kind of inverse phenomena. Particularly in densely inhabited areas, and often as
a result of agricultural development, drainage and processes to achieve water-tightness tend
to dry out aquifers and decrease the level of humidity in the ground, thus accelerating the
dynamics of drainage to the sea. The hypothesis as to the possible effect of these breakdowns
in the continental hydrological cycle upon regional climate changes (drop in average rain-
fall, increased irregularity, rise in risks of flooding, etc.) are currently leading to extremely
interesting studies.
The comments on implied costs due to reservoirs silting up are extremely pertinent but,
should nevertheless be completed with the mention of other costs which are even more relevant;
those of the environmental and economic impacts that this sediment retention produces on
the deltas and beaches along the coastlines. We are now beginning to recognise and value the
importance of the natural flow of these sediments, in order both to compensate for the natural
subsidence suffered by river deltas, as well as the dynamic recharging of beaches and coastal
sediment flows, which are chiefly dependent on river flows. The combination of these impacts,
with the rise in sea level due to the melting polar ice caps, as a consequence of the current
process of global warming, makes the effects of this silting up produced by large dams ever
more extreme and the costs higher.
The progressive drop in the productivity of a large proportion of recently developed
irrigation-based agriculture deserves specific analysis, since this is a worrying tendency and a
direct result of the current development model in which short-term interests override medium
and long-term ones, and kill off any chance of sustainability. The final report by the World
Commission on Dams draws attention to the seriousness of increasing salinisation in many of
these new irrigated lands which at present affects 20% of irrigated surface areas. Furthermore,
the resulting soil degradation usually leads to abandonment of land which would otherwise
have been productive for dry farming. This abandonment is often linked to erosion processes
which, in turn, lead us to irreversible cycles of desertification.
Irrigation efficiency, a key issue: more crops per drop. Observations and comments 121

Special mention must also be made of the direct and indirect impact of drying out large
areas of wetlands worldwide. The argument behind draining and drying out these wetlands is
often based on the need to make them more productive in the fight against hunger and poverty.
Nevertheless, we are gradually becoming more aware of the importance of these ecosystems,
both as generators of biodiversity, and hence of protein (especially fisheries), which is key
to the diet and subsistence of many communities. With the destruction of these wetlands,
their function of regulating and regenerating the quality of water has been upset, seriously
affecting the availability of water resources. Moreover, serious damage has been done to the
rich biodiversity of these ecosystems.
Very often many of these sources of food (fishing, hunting, forestry and small-scale agricul-
ture etc.), closely linked to the sustainability and good health of ecosystems, are not accounted
for in official statistics, since they do not form part of the usual market structure. Neverthe-
less, they have enormous importance in the real economy of many communities, especially in
impoverished or developing countries.

2 TRENDS, PARADOXES AND SOCIO-ECONOMIC CONTRADICTIONS

Prof. Frederick’s arguments regarding the economic problems which large, new hydraulic
projects all over the world have come up against are both interesting and accurate. His treat-
ment of the relentless effects of the law of rising marginal costs, particularly in relation to large
dams, is similar to that of the final report of the World Commission on Dams (WCD, 2000).
Nevertheless, we must also take into account the complementary effect of the law of falling
marginal profits, which can be seen clearly in new irrigation projects (poorer quality of irrig-
able lands, higher altitudes and/or poorer geo-climatic conditions, etc.). The combined effect of
the two laws leads to frequent negative cost-benefit balances, especially in hydraulic projects,
the chief aim of which is to develop new irrigation. It must also be pointed out that the system-
atic loss of agriculture’s relative profitability has contributed to this situation on a worldwide
level, as most countries have seen that rises in crop prices are well below inflation rates.
Furthermore, rationalisation of water management seems necessary, especially as regards
its use in irrigation, creating incentives which would promote saving and the modernisation
of infrastructure. I must state here most categorically that such rationalisation does not neces-
sarily have to be managed by market dynamics, however, as that approach is inappropriate
for water management, except in specific cases of strongly regulated markets, such as the
Water Banks in California. Without Professor Frederick’s work specifically defending the free
market as a strategy for economic rationalisation, he does make comments which I feel it
is fitting to question or, at least qualify. Specifically, it seems appropriate to underline that
this rationalisation does not have to be directed by free market patterns which, in general, are
inappropriate for handling the ethical, social and environmental values of water, beyond that
of purely economic wealth. I do, however, believe in the advisability of promoting (strictly
regulated) public markets for water rights, such as the California Water Banks, in order to
improve the management of business-water in periods of drought (Arrojo & Naredo, 1997).
Subsidies for irrigation water in both developed and developing countries have clearly
negative implications for efficient water use. However, the option of abandoning such subsidies
raises fears about the economic viability of many family farms and the impact this would have
on food prices.
122 P. Arrojo

The worldwide contrast between experiences of irrigation with subsidised surface water
and – generally – unsubsidised groundwater highlights the pressing need to review this kind of
subsidy. In fact, most of this groundwater irrigation, following the criteria of full cost recovery,
is not only viable but much more efficient and profitable than most irrigation using surface and
subsidised water. This review must be undertaken on a case-by-case basis, guided by caution
at the decision-making level, and consistent with the paradigm of sustainable development.
In the most extreme cases of impoverished farmers, direct subsidies to the farmer would be
preferable to general water subsidies. That type of direct, equivalent subsidy would avoid
the financial impact on farmers, while at the same time encouraging them to save water by
improving efficiency.
When applying economic rationalisation measures and processes to modernise irrigation,
no generalisations should be made. Each case should be studied individually, from both the
technical and ecological point of view, and particularly from the social one. Many traditional
irrigation methods provide overall efficiency (including river ecosystems), albeit with scarce
technical effectiveness in terms of running. In addition, many of these traditional irrigation
systems constitute basic sources of subsistence for many communities, communally managed
and highly socially effective, despite the fact that they clearly do not form part of the market’s
global dynamics. In these cases, irrigation flows cannot be considered as business-water but
rather as life-water for these communities. To attempt to impose modernisation processes (both
technical and economic) in these cases may be not only unwise but also counterproductive.
With respect to the increase in food prices, this could have positive rather than negative
consequences. This price increase would be an incentive for sustainability, as it would reflect
the real value of water scarcity. In so far as these possible subsidies would reflect the real value
of water, not only as regards recovering investments, energy and running costs, but also the
scarcity value of water itself and the resulting environmental costs, a rise in prices could provide
incentive for a more reasonable and sustainable use of available resources. The key to the
question is who would absorb the repercussions of these costs. Obviously from an international
context such as the one imposed by the World Trade Organization (WTO), in which the poorest
countries and weakest sectors of society are particularly vulnerable to a market in which the
most powerful impose their rules, the impacts on these weakest sectors could be extremely
negative. In this context, measures should be considered which would compensate for and avoid
these negative impacts. Similarly, the marginalized social classes (impoverished consumers)
in urban areas of most countries should be cushioned against these effects.
In the medium term, the solution to poverty must depend upon countries’ own development
and even their own food sovereignty, if political sovereignty is not guaranteed by international
institutions through a worldwide democratic order.
We must not, however, forget the root of the problem if we wish to resolve it:

– Firstly, a large proportion of the problems of poverty are due to the destruction and break-
down of the rural social fabric, further aggravated by the fall in international agricultural
prices as a consequence of massive subsidies to industrial-scale agriculture in the most
developed countries, particularly the USA and the European Union (EU).
– Secondly, the solution to poverty must be based on the development of these impoverished
countries themselves which generally involves consolidation of their primary sector.
– Lastly, in many cases so-called food sovereignty becomes an indisputable objective, partic-
ularly in so far as political sovereignty is not clearly guaranteed by international institutions.
Irrigation efficiency, a key issue: more crops per drop. Observations and comments 123

For all these reasons a true assessment is needed of the rural environment in general and of
agriculture as a livelihood in particular, not only as regards accepting the real costs of irrigation
water but also as regards agricultural production as a whole, including its underlying social
and environmental values.
However, when the discussion centers on food scarcity, international free market scenarios
and low crop prices, a major paradox becomes apparent. Scarcity, huge demand and low prices
without doubt form a strangely paradoxical triangle. If we take a closer look, we shall surely
find profound contradictions, as we shall later explain.
To start with, we have to consider whether the solution to hunger, poverty and inequality
can be successfully dealt with at all by using a free market strategy without environmental,
social, or ethical rules which regulate it.

3 THE ROLE OF THE MARKET IN THE STRUGGLE AGAINST HUNGER


AND FOR SUSTAINABILITY: POSSIBILITIES AND TRAVESTIES

The present irreversible process of globalisation without doubt offers new approaches to
dealing with world problems of hunger and poverty. The trading of virtual water is one
such approach. However, the logic of the free market is riddled with falsehoods and must be
controlled and regulated from ethical and sustainable perspectives.
Free trade could solve many problems of water scarcity by importing products whose
production requires the intensive use of water (mainly agricultural products). In this way,
countries with water scarcity would save their own water resources and put them to use for
products with higher added value (thus increasing economic efficiency in water use).Very
often, however, cutting down on – or even cutting out – staple food crops basic to national
needs might imply dangerous strategic weaknesses, leading to dependence on foreign countries
and thus endangering their security and independence.
Current unilateral international trends, as led by the present USA government, are weak-
ening international institutions such as the United Nations (UN), as well as trust in these
strategies. This unilateral approach and the lack of multilateral guarantees for countries, which
could later face possible trade embargoes, is dissuading them from developing virtual water
trade strategies.
If we look closely at the agricultural market’s liberalisation policies imposed by the WTO,
the aforementioned paradoxes and contradictions reach a scale difficult to understand and
accept, particularly when we consider the protectionism of the most powerful countries
(especially the USA and the EU), based on production and export subsidies.
The resulting combination of this liberalisation (ignoring ethical, social and environmental
values) and the protectionist policies of the great powers is merely worsening the present
problem. It not only fails to deal with the questions of hunger and poverty but is also destroying
the traditional structure of agricultural production, accelerating the abandonment of rural areas,
and encouraging the rapid increase in urban growth, as well as increasing these countries’
dependence on international trade for their food supplies.
The key question is often not how much is produced but rather how it is produced. Traditional
production methods are not usually the most efficient ones from an international market view-
point but are generally extremely efficient in managing values such as equity and sustainability.
As these methods are based on wise, deeply rooted community-based principles, they are often
124 P. Arrojo

highly eco-socially efficient. It is clearly inconsistent, though, to expect international free mar-
kets to be able to manage these values without suitable legislation and regulating institutions.
To present the problem of hunger in the world as a simple physical equation of production
versus needs not only falls short of its complexity but can also even lead to errors. These errors
will become serious if the WTO free market model is taken on board as the way of managing
this equation. This model, far from offering solutions, is in fact aggravating the problems of
hunger and environmental unsustainability in the world from the structural standpoint.
With these comments, I do not, however, intend to deny the value of the technical arguments
presented by Prof. Frederick, who focuses on improving the efficiency of irrigation and
revitalising rain-fed crops on dry lands. Nevertheless, I believe that these changes should
be made within a new international socio-economic context which differs greatly from the
present one, and this is not reflected in the Towards Sustainability scenario proposed by
Professor Frederick.
Furthermore, beyond resolving the serious contradictions implied in the USA and EU’s
protectionist agricultural policies, we must question whether or not merely allowing developing
countries fairer opportunities to sell their agricultural products on the world market would
automatically lead to better conditions for the majority of rural poor. The general answer must
be NO, unless several important questions, such as the following, are clarified first:
(1) Who currently has or will have access to international markets in developing countries
in the future? In most cases, only the richest farmers and large companies reach these
complex markets.
(2) What are the political conditions of these countries? Democracy, human rights, public
information and protection for the social and environmental rights of communities.
(3) The impact on rural areas when industrialised agriculture is imposed, taking into account
the destruction of traditional agriculture and the social fabric, leading to fragile and
unbalanced development with a dangerous dependence on single crops.
Following free market principles, the WTO has left the social and environmental aspects
of production to one side. This gives an advantage to large-scale farmers and companies and
threatens traditional livelihoods and the rights of communities.
Hence, it is imperative that the WTO should recognise that social and environmental values
are safeguarded by traditional production methods, or other modern techniques based on
ecologically friendly agricultural methods. It is essential to set down regulations to protect
these livelihoods and forms of socio-economic organisation, which are so highly efficient in
reaching the goals of eco-social sustainability, and are so often formally proposed and accepted
in international forums. International institutions such as the UN must be strengthened, so as
to guarantee basic human rights such as the right to food, health or drinking water, over and
above the rules of the free market or, better still, to guarantee these rights as regulations and
limits for the free market.
International agricultural prices should be increased, as is shown in the sustainable scen-
ario presented by Prof. Frederick. This would mean following the laws of the market, whereby
what is scarce but necessary becomes more expensive. This market must, however, be suitably
guided by ethical, social and environmental objectives (eco-social sustainability), as has been
the case in the developed countries’ approach to attaining a high level of social welfare. Food
production must become more profitable, in the general context of a worldwide economy, thus
bringing income improvements essentially for small scale producers and family farms. This
Irrigation efficiency, a key issue: more crops per drop. Observations and comments 125

would be an important impulse for the rural economy, increasing food production and facili-
tating efficient modernisation processes. Defending the opposite – i.e. the present situation
of pitiful prices under pressure from subsidies and export aid for USA and EU farmers as a
strategy to fight against world hunger – is simply sarcasm.

4 ENVIRONMENTAL SUSTAINABILITY AS A KEY FACTOR IN


FOOD PRODUCTION

The problem of the sustainability of ecosystems is often presented as restrictions which hinder
economic development and impede solutions to serious issues such as world hunger. This
analysis is as perverse as it is mistaken. In Professor Frederick’s own documents, he presents
significant data which questions this so-called contradiction between environmental sustain-
ability and economic development, with special reference to the ecological disaster of the
Aral Sea.
Although this case is probably the most widely-known, it must be pointed out that it is
merely one more example of a wide range of problems concerning the impoverishment of
agricultural and biological productivity, as a consequence of unsustainable intensive prac-
tices. These practices are deteriorating the quality of soil, water and other natural resources in
the medium and long-term. Salinisation of lands due to new irrigation is particularly worry-
ing. This problem already affects about 20% of the irrigated land surface (WCD, 2000) and,
coupled with processes of deforestation and the abandonment of traditional agricultural areas,
is opening up the floodgates to erosive phenomenon which irreversibly complete the cycle of
desertification in many regions.
We must also consider the mid and long-term consequences of extensive contamina-
tion caused by modern agricultural techniques based on the widespread use of pesticides
and fertilisers. These impacts, together with drying wetlands and serious effects on deltas and
mangrove swamps, are causing a breakdown in the natural production of protein (fishing and
hunting) at land and sea. These impacts are often the result of complex synergies, which make
precise cause-effect analysis difficult. This leads to evasion of responsibilities. Nevertheless,
the combined result of all the impacts resulting from the multiple breakdown of aquatic and
soil ecosystems is devastating. What is even worse is the way in which this is breaking down
the traditionally stable relationships between many communities and their natural environ-
ment, provoking the destruction of the rural fabric. Perhaps the most serious indirect impact
of the destruction of rural livelihood is the impact of the breakdown in traditional models
of food production (agriculture, forestry, fishing and hunting, etc.), firmly integrated within
these societies and, as previously stated, extremely efficient in the resolution of social and
environmental problems (high eco-social efficiency).
In this sense, the linkage between water and agriculture of which Prof. Frederick speaks,
quoting arguments and conclusions from the World Commission on Water for the 21st Century
(2000), must be characterised as a relationship between agriculture and environmental sus-
tainability involving both the sustainability of aquatic continental and littoral ecosystems and
the sustainability of land. This is, without doubt, the consistent view from which Rosegrant,
Cai, and Cline (quoted by Professor Frederick) attempt to find “an alternative path that is
sustainable and one which will meet growing water and agricultural demands”: the path that
must be conceptualised as eco-social sustainability.
126 P. Arrojo

I would like to express my strong agreement with the re-evaluation and development of rain-
fed agriculture in Professor Frederick’s Toward Sustainability scenario, as being a key factor. It
is essential to do away with the myths surrounding irrigation and recognise the value of rain-fed
agriculture, due to its greater agro-environmental suitability for sustainability scenarios.
With reference to the sustainability of aquifers, although it would often be necessary to
review and lower the scale of water drawn, I have serious doubts that the potential of the
integrated use of ground and surface waters has been correctly quantified. The lack of data
on aquifers and the priority which most governments place on the use of surface waters via
large hydraulic works, make integrated use of surface and groundwater one of the options
with greatest potential for development in many areas.

5 THE NEED FOR A NEW CULTURE AND THE SIGNIFICANCE OF EDUCATION

Certainly, the limited space of a scientific paper does not permit Professor Frederick a detailed
explanation of the methodology used to design the scenarios presented. All the same, I would
still express a certain degree of scepticism regarding the extensive world figures and data
often used in the design of this kind of long-term scenario. These statistics tend particularly
to leave aside both scattered traditional economies, which have not been integrated within
national and international markets, and also underground economies. This problem has much
greater relevance in poor or developing countries, especially if the questions we wish to study
are linked to underdevelopment, poverty or food sufficiency. In these cases, key elements and
data overlooked may play a more decisive role in both the diagnosis and possible solutions.
One of the key issues for the medium and long-term future is undoubtedly the cultural and
educational factor. The world is becoming increasingly more complex and globalised, and
large-scale solutions will rarely be successful if they are not designed and applied through
participation-based processes involving the communities affected in each region.
The very evolution of populations, which is usually presented as an exogenous and inex-
orable element, determining the forecasts of any imaginable future scenario, will depend on
cultural and educational factors which must be tackled. On this point, the educational and
cultural level of women is considered to be a particularly essential factor in this respect.
The change in eating habits, which is currently pushing meat consumption (with high
costs for virtual water) above healthy limits in many countries, may indeed alter agricultural
strategies and their respective water needs.
Amongst the lines of action needed to develop a collective intelligence based on a New
Culture of Water and Sustainability, the need to develop suitable policies for regional and urban
planning stands out as a salient feature. I wonder if the models and scenarios of Rosegrant,
Cai and Cline, quoted by Prof. Frederick, include this factor.
How to produce can and should be as important, or more, as how much to produce, since
eco-social efficiency can, and does, in fact, differ greatly, depending on present and possible
different models and lifestyles.

6 THE ROLE OF TECHNOLOGICAL DEVELOPMENT

Obviously one of the key elements implied in the more crops per drop slogan that stands out
in Prof. Frederick’s work, is to be found in the application of new technologies. Nevertheless,
Irrigation efficiency, a key issue: more crops per drop. Observations and comments 127

I believe it would be useful to emphasise the significance of technological development in


the qualitative dimension, beyond its undoubted importance for the increase in quantitative
efficiency emphasised in this slogan.
The new technologies of semi-permeable membranes, combined with energy recovery tech-
niques in desalinisation plants, are leading to increasingly lower energy costs. These costs are
presently between 0.40 and 0.50 a/m3 , while energy requirements are around 3.5 Kwh/m3
(Arrojo, 2003a). This is revolutionising the question of viable alternatives, especially as
regards supplying urban coastal areas. The long term implications of these technological
advances are even greater if we bear in mind the high quality and reliability of the drinking
water thus obtained. It should be pointed out that already now, and more and more in future
forecasts, the most conflictive situations will be linked to growing urban population density in
coastal areas.
The decreasing cost of these new technologies means that desalinisation can compete
advantageously with the regulation and long-distance transportation of surface water via large-
scale infrastructures, the cost of which shoots up above 1 a/m3 for distances over 500–600 km
(especially if environmental and even security costs are accounted for). This offers alternatives
which reduce the pressure and impacts on continental aquatic ecosystems, as are now being
embarked upon in Spain, after large-scale water diversion from the river Ebro was rejected
and desalinisation of briny and sea water has been made a priority.
This technological development is in tune with the growing priority of the quality factor
rather than quantitative problems, which appear to prevail in Prof. Frederick’s work. Already
today, and in the future, the most pressing problems of water in the world concern quality
more than quantity, although the two aspects are certainly linked.
Avoiding contamination, by both ecological production models and suitable technology
for the treatment of returns will become the priority, surpassing the polluter pays principle.
From this new perspective (which gives pride of place to the preservation of quality), the
conservation of ecosystems and of their natural capacity for regeneration becomes a priority.
This is the logic which tends to be imposed in advanced countries. The possibility of increasing
water reuse must enable much greater availability of resources without increasing the pressure
on ecosystems.
The technologies involving re-use of urban water for agriculture or even for secondary
uses (lavatories, gardens, car-washing, washing machines, etc.), through a double network
(easy to implement in new urban areas) provides raw water of sufficient quality at costs below
0.12 a/m3 (Arrojo, 2003b).
Perhaps the greatest advantage of these new technologies resides in its modularity and
flexibility as opposed to the rigidity of the major traditional hydraulic systems. It is fundamental
that we should fit quality water production to the real needs of each moment using modular
strategies, and once and for all abandon grandiose, oversized public works which are rarely
sufficiently justified and always backed by declarations of somewhat suspect general interest.
Through such modularity and flexibility, greater transparency in cost allocation would be
achieved, bringing with it an easier introduction of demand management strategies.

REFERENCES

Arrojo, P. (2003a). Spanish National Hydrological Plan: reasons for its failure and arguments for the
future. Water International, 28(3): 295–303.
128 P. Arrojo

Arrojo, P. (2003b). El Plan Hidrológico Nacional: una cita frustrada con la historia. RBA Editores-
Integral, Barcelona, Spain.
Arrojo, P. & Naredo, J.M. (1997). La gestión del agua en España y California. Colection Nueva Cultura
del Agua, 3. Editorial Bakeaz, Bilbao, Spain: 192 pp.
WCD (World Commission on Dams) (2000). Dams and development. A new framework for decision-
making. Earthscan Publications Ltd., London and Sterling, UK and USA.
World Commission on Water for the 21st Century (2000). A Report of The World Commission on Water
for the 21st Century. Water International, 25(2), June: 284–302.
IV

Virtual Water
CHAPTER 8

Virtual water – Part of an invisible synergy that ameliorates


water scarcity

J.A. Allan
School of Oriental and African Studies (SOAS), London University, UK

ABSTRACT: The purpose of the chapter will be to demonstrate that there are a number of economic
processes that have the capacity to ameliorate local water scarcity. A number of arid and semi-arid regions
and economies encountered water scarcity in the past thirty years. Many more will encounter water
scarcity in the next three decades. The analysis will review briefly a threefold synergy of ameliorating
processes – first, the contribution of unaccounted water in soil profiles [known as effective rainfall to
hydrologists and engineers] to the production of food staples; secondly, the global role of virtual water
in ameliorating water scarcity in dry regions; and thirdly, the impact of socio-economic development on
water management options. All three processes have the characteristics of being economically invisible
and politically silent. But they ameliorate conflict in the easily politicised domain of water allocation
and management. Their impacts are determining with respect to solving local water deficits. The role of
a fourth ameliorating technology, desalination, will also be examined.
The status of the world’s water resources will also be reviewed and estimates of the proportion of
global water resources, which become involved in virtual water transactions will be provided. The main
focus of the chapter will be on the extent to which virtual water in trade has successfully met the past
and current needs of water deficit regions as part of the threefold synergy – soil water, virtual water and
socio-economic development. There will be a very preliminary evaluation of the role of virtual water in
meeting the future needs of water scarce regions during the demographic transition of the 21st century.
It will be shown that the water, food and trade nexus is not easy to model because of the dynamics of
the political economies in the North and the South. Water sector policy-making is subject to evolving dis-
courses, which can easily de-emphasise the underlying environmental and economic fundamentals. The
chapter will conclude that invisible and silent virtual water in the invisible threefold synergy will provide
the big water, that is the food-water for water scarce regions. Its invisibility and silence will, however,
have the effect of attenuating the pace of water policy reform. Both reforms addressing water use effi-
ciency and those giving consideration to the environmental services provided by water will be attenuated.
A range of demographic, economic, social and political theory will be used to frame the discussion.

Keywords: Virtual water, soil water, socio-economic development, the threefold synergy, global
water-food-trade nexus, incidental benefits, water policy, political economy

Not all waters are equal: some are more evident than others.

1 INTRODUCTION: SOLUTIONS TO CLOSURE AND WATER SCARCITY

The title of the programme/book to which this chapter is a contribution asks the question
is there a current or a future global water crisis? Is there enough freshwater – surface and
132 J.A. Allan

groundwater, and soil water to meet the demands of current and future populations? Within
the question there is an implied additional question – are some regions and some river basins
facing current water crisis and will they face irreversible water crises?
The purpose of the chapter will be to show that there is sufficient freshwater and soil water
in the world to meet current and future water needs if we accept the demographic predictions
in currency since the UN World Population Conference in Cairo in 1994. The low estimates
of the world’s population in the late 21st century are about 8000 millions. High estimates
are about 11,000 millions. The present world population is about 6500 millions requiring
about 6500 km3 /yr of freshwater and soil water. The future population estimates imply annual
requirements of between 8000 and 11,000 km3 of freshwater and soil water.
The conceptual and analytical contribution of the chapter, in a review of the future adequacy
of global water – the subject of the book – is to show that trade in commodities has already
achieved a pivotal position in enabling the comparative advantages and disadvantages of
regional water endowments to be balanced. A commodity such as grain, one of the major
traded commodities, requires for its production amounts of water 1000 times the weight of the
commodity. An economy receiving a tonne of grain does not have to face the economic, and
more important, the political stress of mobilising 1000 tonnes (1000 m3 ) of water. The future
role of virtual water in remedying the increased regional water deficits will be proportionately
more important as populations increase in water deficit regions.
Pessimists such as Lester Brown (2003) and Sandra Postel (1999), and Postel & Richter
(2003) argue that the world does not have the water resources to meet the needs of future
populations. They point to the strains facing the water scarce Middle East and especially the
situation in heavily populated South Asia and China where water quantity and quality are
worsening problems and groundwater resources are under severe pressure.
Optimists such as the author of this chapter (Allan, 2001), and most of the international
agency staff and many scientists intimate with demographic and water resource problems,
(World Bank, 2003; IFPRI, 1995, 1997; Dyson, 1996; Rosegrant et al., 1995; Rosegrant &
Cai, 2002; Brichieri-Colombi, 2004) judge the challenge of providing enough food to meet
global needs as addressable, albeit with problems in sub-Saharan Africa. These problems are
attributable to poverty rather than to water poverty.
It will be shown that one of the important consequences of the existence of the trade
associated with the concept of virtual water is to make it possible for water users, water policy-
makers and the general public to embrace simultaneously optimism and the pessimism. It has
been noted that “pessimists are wrong but useful; optimists are right but dangerous” (Allan,
2001). Optimists and pessimists interpret the underlying fundamentals differently. In practice
the underlying environmental and economic fundamentals are rarely effective starting points
from which to campaign for changes in water policy. Governments and peoples in severely
water scarce regions are in denial about their water resources and their policy options.
An understanding of the politics and local history of water policy-making processes is
a much better starting point for a water policy reform campaign. This is not a new idea.
Marx pointed out the determining role of politics. The historically evolved abstract domain
of politics overwhelms the information coming from science and sub-optimal markets in
what Marx termed the concrete economic domain (Fine & Saad-Filho, 2003). The concrete
provides the foundations for the (political economy) superstructure, where those with power
make allocative decisions. Evidence on the concrete is only very selectively included in policy-
making discourses. Scientists rarely draw up or impact policy-making agendas. Nor are they
Virtual water – Part of an invisible synergy that ameliorates water scarcity 133

normally much involved in developing politically controversial allocative policy. They can
help to raise awareness of how things are. But they normally have to struggle mightily to bring
about significant shifts in perception.
One of the reasons that water science and water professionals have limited purchase on
policy-making is because there are economic processes outside the water sector and its water-
sheds, which solve water scarcity problems. The analytical tools of hydrologists, environmental
scientists, and hydraulic engineers are not effective in these problemsheds which lie beyond
the watersheds of the water sector. The agenda of the meeting to which this chapter was a
contribution is evidence of the norm amongst water professionals and scientists that water
problems are mainly viewed to be the concern of professionals and scientists associated with
water and the productivity and quality of water.
The main reason that water professionals have limited influence on water policy discourses
is their limited experience of advocating and implementing water policy reform in the late 20th
century. Before the 1970s water scarcity was little in evidence. Technical, supply management
measures were economically appropriate and had not yet been identified as environmentally
hazardous. By the 1980s the negative impact of supply management water policies on the
environment required that water professionals tell the truth to power. Technical remedies were
no longer always appropriate. By the 1990s the message that water should be valued required
that politically costly re-allocative measures be part of the unwelcome truth to be digested
by the powerful. In practice water scientists and professionals did not tell truth to power.
That controversial task was undertaken by environmental civil movement groups. Many water
scientists fell in behind the campaign and have reinforced the environmental project.
After two or more decades of reorienting their own approaches to managing water scientists
and professionals are still undecided about how to cope with the politics of water policy-
making. For centuries, even millennia, water professionals solved the water problems of the
societies managed by political elites. They did this without much consideration for the value of
the environmental services of water or of the economic value of the water inputs to society and
agriculture. Recognizing these two issues as new fundamentals was challenging. Conveying
their unavoidable importance to water users existing in rural poverty and to the politicians
running their economies has proved to be very difficult. Water users with few or no alternative
livelihood options to irrigated farming, and water policy makers wanting to avoid paying
high political prices associated with re-allocative reforms cannot manage water demand. The
outcome was to take water policy discourse into realms of painful uncertainty.
Water professionals had little experience of a world where problem solving was social
and political rather than technical. The skills needed to operate in the political domain are
unfamiliar. They require the capacity to deal with ambiguity and uncertainty. In another
policy domain, Teller, the nuclear physicist, had no such scruples and knew how to influence
politics. “He knew that it was more like magic than logic, and in spinning his spells, he
was both dishonest, wasteful, and at times dangerous. He never saw this as evil or immoral.
He once corrected Oppenheimer’s famous comments that ‘scientists have known sin’. As far
as Teller was concerned ‘scientists have known power’ and what it demands.” (Goodchild,
2004).
Water scientists and water professionals are still unsure whether they should know power,
and remain uncomfortable telling new truths to power. Activists from the green movement
have been much more successful in shifting the debate and water policies than the science
community and water professionals. The activists are regarded by many water professionals
134 J.A. Allan

and scientists as dishonest, spinners of spells and at times dangerous. But they have influenced
the discourse and associated policy reform. They have not achieved their goals but they have
had a very significant discursive impact on the ideas in currency and especially on the valuation
of the water services in the environment.

2 SOIL WATER AND VIRTUAL WATER: BRIEF DEFINITIONS OF TWO


NON-EVIDENT WATER SECTOR PROCESSES

“Things do not always appear as they are.” (Fine & Saad-Filho, 2003: 4, quoting Feuerbach &
Marx).

The reason that food production and water have been emphasised thus far is because water for
food is the dominant use of water. About 90% of the water needed by an individual is devoted
to food production. The remaining 10% is needed for drinking, domestic use and the support
of non-agricultural livelihoods.
The three departures from conventional approaches to water allocation and management
integral to the message of this chapter are the following. The first generally ignored feature,
though hydrologically and economically fundamental, is soil water. The second is virtual
water. Virtual water is a new concept to the water sector. But the term is merely an example
of Ricardo’s powerful, two century old, notion of comparative advantage. Wichelns (2004)
has added a purist economics definition insisting that virtual water is an example of absolute
advantage because it does not have embedded in it optimal production and trading strategies.
He recognizes that “the virtual water metaphor addresses resource endowments, but it does not
address production technologies or opportunity costs. Hence, the metaphor is not analogous
to the concept of comparative advantage. The metaphor can be helpful in motivating public
officials to consider policies that will encourage improvements in the use of scarce resources,
but comparative advantages must be evaluated to determine optimal production and trading
strategies”.
Prior to human consumptive water-use natural water resources supported only biological
and hydrological systems. All the consumptive use of water by human populations has to be
taken from the diverse environmental water resources available to their political economies.
Available water includes surface waters, groundwater, soil water and atmospheric water. Sur-
face waters, groundwater and soil water are the waters that the human population use for
their economic, social and amenity purposes. Surface waters and groundwaters are taken into
account by water scientists and water policy-makers insofar as they can be quantified. Soil
water is generally ignored. Things do not always appear as they are.

2.1 Soil water: the first invisible


Soil water is not accounted in national water budgets. The Shiklomanov (2000) research group
is an honourable exception in that soil water appears in all its datasets. Nor does soil water
figure much in the statistics of the international agencies. These datasets focus on freshwater
at the surface and in the aquifers (FAO, 2003; FAO, 2004 – Aquastat; World Bank-WRI, 2003).
The Aquastat database does include a reference to the rainfall that contributes to the generation
of national water resources. But it is not possible to derive an estimate of the soil water element
of a national budget.
Virtual water – Part of an invisible synergy that ameliorates water scarcity 135

This discussion will not emphasise the water services provided by the water in the natural
environment. The absence of such discussion does not mean that such water services are
regarded as unimportant. There is not space here to address the topic adequately.

2.2 Virtual water: the second invisible


Virtual water has been conceptualised and defined elsewhere (Allan, 2003). The concept has
also been critiqued and debated (Merrett, 2003). Virtual water is the water associated with
the production of commodities. Research has focused on the water required to produce grain
staples such as wheat. It requires 1000 m3 of water (1000 tonnes) of water to produce a tonne of
wheat. Importing a tonne of wheat has a very favourable impact on the economies that endure
water deficits. It follows that virtual water should have a place in reviews of global and regional
water resources and in strategizing water policies. The water sector and the freshwater in a
nation’s rivers and groundwaters are not a sufficiently comprehensive basis for quantifying,
analyzing and optimizing the allocation and management of water resources.
To see things as they are it is necessary to review water resources and water transferring
processes comprehensively. Water allocation and management is a global issue and a global
challenge. The same is true of many natural resources, for example oil. The global hydro-
economic system can solve the problems of water scarce regions. Non-hydrological systems,
such as trade, are as problem solving as the water and water flows within watersheds captured
by hydrology. The proportion of the water re-distribution challenge successfully addressed
by virtual water processes is already impressive. It has the potential to be of even greater
significance. Hoekstra & Hung (2002: 25) estimate that 695 km3 of freshwater and soil water
entered international trade in virtual form out of the 5400 km3 of water used to produce crops
in the 1995–1999 period.
Livestock products are particularly water intensive. A study of the virtual water content
of the 1995–1999 trade in livestock products (Chapagain & Hoekstra, 2003a, 2003b, 2003c)
calculated the figure to 336 km3 /yr. Most livestock products are raised on soil water. But a
proportion of such products are produced from irrigated fodder. By adding the annual global
virtual water figures for crop production − 695 km3 – to the 336 km3 for livestock products,
we get a total global figure of 1031 km3 /yr. There is some double accounting in this number
but the calculation nevertheless does confirm the very important part played by virtual water,
mainly derived from soil water, that enables sustainable water resources management across
the globe.
All the volumes of water discussed above are situated in rising long term trends, although
there is a current temporary leveling off in the trade in major grains. The leveling off is the
result of improved efficiency as measured by returns to water in both the Northern and the
Southern economies (Dyson, 1996, 2001).
Meanwhile estimates of current global freshwater use in irrigated farming suggest that the
total mobilized is about 1430 km3 /yr – that is 26% of the total water used in crop production
(Rosegrant & Cai, 2002). The same authors suggest that by 2025 the use of irrigation water
could rise to 1480 km3 /yr. Clearly the virtual water solution, associated with 13% of global
water used in crop production is significant as a problem solving process. The trade in livestock
products adds to this global percentage. Virtual water is especially significant in addressing
those problems, which local technological solutions cannot address in water scarce regions.
Just as important as its proportion of total water use in crop production and its capacity to
136 J.A. Allan

solve otherwise un-addressable problems in water scarce regions, is its flexibility. This quality
of flexibility will be discussed below.

2.3 The concept virtual water


Virtual water is a powerful concept as well as a powerful if invisible operational solution, in
that, like comparative advantage, it is both an intensive and an extensive concept (Weber, 1904,
1917). Such concepts help us communicate, even if an intensive/extensive notion is at first
unwelcome to water policy-makers and especially to politicians riding the politicized water
sector tiger. First, in the case of water, the concept links the intensive process of combining
water in for example the agronomic and technical processes of food crop production. Secondly,
it is an extensive concept that it links water with the transactions that move and trade water
intensive commodities such as grain. Without such a concept the activities of production and
trade have to be conceptualised separately as recommended by Merrett (2003). Happily the
concept of comparative advantage is a useful precedent in commending the integration of the
intensive and the extensive processes to professional economists.
This intellectual inspiration of comparative advantage led to the Heckscher-Ohlin (H-O)
Theorem of trading factors of production (Hakimian, 2003). The notion is something of a
mega-intensive/extensive trade theory. The theory has intuitive relevance to water, as water
can readily be defined as a factor of production. Although its role in economic systems can
be very difficult to track. Unfortunately attempts to deploy the H-O Theorem have at best met
with mixed success. The absence of success is partly the result of inadequate datasets. Even
more important are other factors than resource endowments which impact directions and levels
of trade. For example economies of scale, trade policies and demand factors can be important
(Bowen et al., 1987; Krueger, 1977; Harkness, 1978, review extensions and empirical tests;
also Lawler & Seddighi, 2001; Krugman & Obstfeld, 1991).
One of the outcomes for the author of looking at the literature on trade in factors of
production is to deter too much commitment to quantifying virtual water processes. The work
of both Earle (2001) and Hakimian (2003) shows that while the metaphor remains convin-
cing it is extremely difficult to emerge with persuasive numbers when attempting empirical
confirmation.
The same is true, but to a lesser extent, of the preliminary work on water by the scientists
at the Institute of Hydraulic Engineering (IHE) in Delft (Hoekstra & Hung, 2002). Their
research on virtual water came up with global numbers that fit quite well with what specialists
expect to see. 5400 km3 /yr of water are estimated to be used in crop production according to
the heroic use by Hoekstra & Hung (2002) of the best available global and national datasets.
The figure makes sense in relation to the estimates of global population of say 6500 millions
at 2000. A rough estimate of annual domestic and non-agricultural livelihood water for the
global population would account for a further 650 km3 . The food/agricultural water plus
the domestic/non-agricultural livelihoods water would then total 6050 km3 /yr. This total is
remarkably close to the number that a quick and dirty estimate of global water consumption
based on a consumption of 1000 m3 /yr per person.
The Hoekstra and Hung work is of particular value because it has provided numbers that
confirm the big water/small water idea. They show that big water, the water used to produce
food and other crops, accounts for about 90% of the water used by the human population.
The number remains counter intuitive for rural communities where populations are rising. It
Virtual water – Part of an invisible synergy that ameliorates water scarcity 137

is especially challenging to rural societies living in poverty where non-agricultural livelihood


options are absent. The idea is seriously unacceptable to politicians managing economies
where creating more jobs per drop is difficult.
The national numbers can only be as good as the datasets on which they are based and these
are not yet reliable. The IHE work has been extremely useful, however, in that it has calculated
a very much needed first approximation of global and national virtual water transactions. Just
as important as the quantification is the interest it has attracted in a diverse range of water
scientists and professionals in all parts of the world. That interest has turned into a significant
epistemic event, reflecting a new level of engagement by scientists and professionals involved
in water science and water policy-making in both the North and gradually in the South (see
Hoekstra, 2003; Chapagain & Hoekstra, 2003a, 2003b, 2003c; Haddadin, 2003; Zimmer &
Renault, 2003; El-Fadel & Maroun, 2003).
A very important aspect of the both the productive water related transactions and the trading
transactions discussed in this section is that it is soil moisture rather than freshwater that is
used to produce the bulk of the food that is traded (Hoekstra & Hung, 2002). It is soil water,
mainly in humid temperate zones, that ameliorates the water deficits of regions enduring
freshwater and/or soil water deficits, in the arid and semi-arid tropics. Farmers and traders can
mobilize a process with a global reach to achieve a remedy, of which hydraulic engineers cannot
dream.

2.4 The value of virtual water


The issue of the value of virtual water has been usefully discussed by Renault (2003). He
pointed out that water can be differently perceived and differently valued. The perception of
the value of water in economies that export water intensive commodities can be very different
from that perceived in economies that import water intensive commodities. The opportunity
value of the water might be close to zero in a soil water rich environment such as the temperate
regions of North America and Europe. The equivalent embedded water in the water intensive
commodities imported into the water scarce economies of the Middle East would be much
more valuable. In some circumstances the marginal water could be desalinated water with
value of between US$ 0.50 and US$ 1.00. The valuation of virtual water is an important issue
but there is no space to discuss it here beyond drawing attention to the importance of the
economic context in which such valuation is made.

3 SOCIO-ECONOMIC DEVELOPMENT: A THIRD INVISIBLE SOLUTION FOR


THE WATER SCARCE

Distant soil water and virtual water trading processes have, until recently, not been recognized
as being significant in addressing the problems of the water scarce. They have not been a part of
explanations of how things are for the water scarce in for example the Middle East region. The
concepts remain very much part of a minority outsider discourse even in the Middle Eastern
epistemic science and policy communities associated with water allocation and management.
The third non-evident process that contributes to the amelioration of water deficits is socio-
economic development. The achievement of more jobs per drop rather than more crops per
drop has been key to the successful amelioration of water scarcity except in sub-Saharan
138 J.A. Allan

Africa. This ameliorative process is not even on the analytical agenda of most water scientists
and water policy-makers in the South.
Socio-economic development associated with the strengthening and diversification of
economies has achieved impressive levels of allocative efficiency in association with virtual
water related processes. Overpopulated economies and overpopulated regions of some national
economies in dry regions have become diverse and prosperous despite being extremely water
short. The water poverty of a region has not determined that a region would endure poverty.
The very desirable allocative efficiencies achieved in the water sector in the past half century
have, however, been incidental benefits resulting from the socio-economic development.
These allocative water efficiencies were not part of water policies. They were hidden-
hand (after Adam Smith) processes. The hidden hand enabled the perversely hydrocentrically
(Brichieri-Colombi, 2003) inclined water policy-makers and their water professionals to
achieve sustainable economic outcomes. Job creation in industry, services and the public sector
silently and invisibly reallocated water at a scale beleaguered politicians yearn for. Job creation
associated with economic diversification has proved to be a mighty demand management tool.
In addition job creation outside agriculture has generated incomes and revenues with which
political economies have been able to access virtual water in the global system. Economic trans-
formation achieved in the East Asian economies, and in for example Israel (Allan, 2001: 249),
demonstrate how economies can achieve water security through economic diversification.
Without economic diversification there would be no trade in the commodities associated
with virtual water. Without trade in these commodities there would be no demand for the
soil water in water surplus regions. Economic diversification enables a water scarce economy
to articulate its demand for water in a system that can provide it, namely the trade in water
intensive commodities.
Virtual water transactions and diversified job creation have enabled five decades of rela-
tively conflict free global co-evolutionary transition from water abundance to water scarcity.
For the past 25 years potential regional water scarcity has been managed within sustainable
regional and global water management regimes. The very effective invisible threefold synergy
of soil water access, virtual water processes and socio-economic development has provided a
very effective and a politically stress free regime.

4 THE TIMELY AVAILABILITY OF THE FLEXIBLE THREEFOLD SYNERGY:


SOIL WATER, VIRTUAL WATER AND DEVELOPMENT

The worsening water scarcity experienced in some regions of the world in the second half
of the 20th century was encountered at a fortuitous moment in world economic history. The
first good fortune of the water scarce was the availability of a comprehensive global trading
system in food. This is not to say that the food trade is new. The system has been operating
effectively for millennia. Rome in antiquity was fed by grain and other commodities raised
in North Africa. Comparative advantage is not static. As Wichelns (2004) argues in each case
it is related to the production systems and the opportunity costs that obtain. By the late 20th
century economies such as Egypt, the main North African food exporter in antiquity, were
importing nearly half their food. Two thousand years ago Egyptian agricultural exports to
Italy were substantial. Semi-arid economies facing water scarcity encountered their scarcities
at a moment in history when, fortuitously, the world’s temperate regions had become highly
Virtual water – Part of an invisible synergy that ameliorates water scarcity 139

industrialized. They produced substantial crop surpluses throughout the second half of the
20th century.
A second fortuitous, and invisible, non-water sector, process assisting the water scarce
since the middle of the 20th century has been the economically perverse agricultural policies
of the European Union (EU) and the USA. Their production and export subsidies on wheat for
example, ensured that half-cost grain was available on the world market. The pivotal global
grain importers, the Middle East and North Africa economies and Japan, were both well
able to purchase imports. They have been significantly advantaged by the EU/USA subsidies.
Together they accounted for about 150,000 Mm3 of virtual water in crops traded in the global
system in the last five years of the 20th century. These major importers accounted for 23%
of the total. The author is at the same time aware of the negative economic impacts of low
world prices for grain on economies with very low levels of Gross Domestic Product (GDP) per
head, especially those in sub-Saharan Africa. These economies enjoy no potential comparative
advantage except in crop production. Their farmers are seriously impacted when they cannot
obtain a local market price for their crops because of imports of subsidised staples.

4.1 Variable water demands: the threefold synergy is also flexible and responsive
Virtual water is part of an important threefold synergy – soil water, virtual water and socio-
economic development – that enables communities, nations and river basins to access sufficient
water to meet their variable water needs. A further very important feature of the threefold
synergy is its unmatched flexibility. Since the 1970s it has enabled the extension of a form
of supply management miracle across the globe. The synergy has been capable of meeting
varying regional and emergency food/water demands. Variable demand is normally associated
with drought and its consequences. The threefold synergy can successfully augment policies
that in themselves cannot achieve local sustainable water allocation or address water scarcity
emergencies.
For at least three decades the threefold synergy has proved its capacity to respond to systemic
trends associated with increased demands associated with rising populations – for example
in the Middle East. The system has coped with the closure of the Middle East and North
African water systems and the resulting increased demands for food imports. It has also coped
with the variable demands coming from the former Soviet Union and the Russian Federation.
The biggest variable demands have come from China. These demands have occurred despite
China’s own extraordinary increases in production since 1961. The fluctuating demands for
grain imports from China have also been accommodated. That there is a mix of local and
international solutions is revealed in the decline in volumes of grain entering international
trade in the last five years of the 20th century. Despite the progressive increases in the world’s
population the volumes of grain traded have been declining since the late 1990s.

5 THE THREEFOLD SYNERGY AND THE ENVIRONMENT

A further benefit of the threefold synergy is its amelioration of arid and semi-arid regions’
competition for water between agricultural use and environmental services. By reducing the
pressure on water resources in water scarce regions the threefold synergy has alleviated, or at
least it has had the potential to alleviate, progressively higher demands on the freshwater and
the soil water in the environments of the water scarce. The availability of the threefold synergy
140 J.A. Allan

has enabled demands on local surface and ground waters to level off and in some cases to be
reduced. This is especially the case in Northern semi-arid economies – for example in Israel.
Water has been returned to the environment in Northern semi-arid economies such as those of
the USA western states to reinstate, to some extent, the environmental services provided by
natural hydrological systems (Allan, 2001: 146–148). In the South, the availability of virtual
water has not yet had the same impact.
The shift towards precautionary and green water policies in the North has not been a
response to the economic processes that made it possible. The advocacy of the green movement
has been the important factor in achieving environmentally sensitive water policy-reform in the
North. In this Northern discourse the new resource managing circumstances afforded by the
threefold synergy has enabled pressure on the water in the environment to be alleviated in
these Northern political economies.
The green social movement is poorly developed, however, in most Southern political
economies. In Southern semi-arid economies the apparent amelioration of scarcity by the
threefold synergy can play a negative, rather than a positive, role in strategising the allocation
of water vis-à-vis the environment. Old practices and policies that are economically nor envir-
onmentally appropriate can be left in place because to reform them would only be achieved
at unacceptably high political prices. Paying such prices can be avoided in the apparent water
security provided by the invisible threefold synergy. Water abstraction practices associated
with the livelihoods of poor farmers with no alternative job options tend to remain in place,
including those, which damage the water environment (Allan & Olmsted, 2003).

6 CONCEPTUALISING THE SOIL WATER, TRADE AND TECHNOLOGY


ELEMENTS OF THE THREEFOLD SYNERGY

Figure 1 is helpful in conceptualizing a number of essential ideas that could usefully be


adopted by the water science community as well as by the diverse epistemic group associated
with making and advising on water policy.
The first idea is that of big water and small water. The big water is the 90% of freshwater
and soil water needed to produce the food needs of human populations. Small water is the
10% of freshwater – it can only be freshwater – needed to provide drinking, domestic and
municipal water as well as the water needed for non-agricultural livelihoods.
Secondly, Figure 1 is useful in conceptualising the role of soil water. Two political economies
with similar populations – Egypt and the United Kingdom – located in very different envi-
ronments, can be compared. The diagram indicates the significance of their respective water
endowments. Egypt has to use almost all its average annual availability of between 55,000
and 60,000 Mm3 of freshwater in the Nile River. The figure includes reuse. It has negligible
soil water. Egypt accesses virtual water to meet its rising water deficit. The UK uses about
15,000 Mm3 of surface and ground waters. Soil water is not negligible in the UK water budget.
About 25,000 Mm3 of local soil water and access to virtual water enables the UK to meet its
food-water needs. In both cases the virtual water calculus is a net figure in that both economies
export food commodities as well as importing them.
The calculated water footprints of both economies are quantitatively very similar and nega-
tive (Hoekstra & Hung, 2002). The concept of the water-footprint is a development of the much
more complex concept of the ecological footprint and captures the extent to which a political
Virtual water – Part of an invisible synergy that ameliorates water scarcity 141

economy needs to access resources beyond its boundaries or alternatively the circumstance
where an economy can export its natural resources (Wakemagel & Rees, 1996).
In addition Figure 1 is useful in conceptualizing and highlighting the role of the two
important solutions for any economy entering water deficit circumstances. First, desalinated
water, and secondly, virtual water. The ability to access both of these solutions depends on
the performance of the economy. Both the production of manufactured water and access to
virtual water depend on the capacity to mobilize financial resources. In the case of water
manufacture it is necessary to service the capital investment and the operating costs of

Technological
remedy to
Manufactured desalinated water small national
water deficits

Water-drinking, D
10% domestic, industry and services

UAE
Fresh water at Dependency
the surface & c77%
in groundwater
Water aquifers
used in
agriculture, D
including
fresh-water Total water
use in UK self-sufficiency
irrigation Dependency
systems c35%
Egypt
Dependency
c22%

D
Soil water in the
soil profile
accessed by crops
D
90%

Economic remedy –
the import of
Virtual water remedy to the large food enables
water deficits associated with insufficient water dependent
water for crop production. economies
Very flexible; also invisible & silent to remedy their
large water deficits
thro’ trade
UAE: United Arab Emirates
D: water dependency

Figure 1. (1) The extent of the water dependency (D) of different water deficit economies (Hoekstra
& Hung, 2002: 55–59). (2) How local water can come from freshwater sources and soil water sources.
(3) How manufactured water and virtual water can remedy the water deficits.
142 J.A. Allan

desalination technologies. In the case of the virtual water solution it is necessary to have
the capacity to purchase commodities such as food on the world market.
Desalination is a topic, which deserves a separate discussion. During the past decade one
of the desalination technologies – reverse osmosis – has experienced a dramatic reduction in
operating costs as a result of the fall in the cost of the filtering membranes. Unfortunately
the capital, and especially the operating, costs of both multi-stage flash and reverse osmosis
systems are notoriously difficult to estimate (Stauffer, 2004). The problem arises because
the cost effectiveness of combined power generation and desalination plants depends on the
seasonal interaction of the load factors in the power and desalination domains.

7 VIRTUAL WATER IN RELATION TO GLOBAL FRESHWATER RESOURCES


AND THEIR MANAGEMENT

It has been pointed out above that virtual water moved by trade accounts for about 20% of the
water used in global crop production. Meanwhile most water policy is made on the assumption
that freshwater – being the only water that engineers and farmers can pump and economists
can account for – is the only water that should enter into the planning and policy-making
processes. The Russian hydrologists (Shiklomanov, 2000) and the Chinese water resources
professionals [personal observation] do recognize the role of water in both the hydrological
and the economic cycles. Shiklomanov (2000) suggest that about 5500 Mm3 of water are
involved in agricultural production. He includes production from soil water.
The 1031 km3 of the water used to produce the crop and livestock commodities that
enter world trade account for 19% of the agricultural water use. Research at the Value of
Water Research Unit at the IHE in The Netherlands has provided first approximations of the
virtual water associated with crops (Hoekstra & Hung, 2002) and with livestock products
(Chapagain & Hoekstra, 2003a, 2003b, 2003c) that enter world trade. The IHE research esti-
mates that 695 km3 of water are associated with the crop commodities traded (Hoekstra &
Hung, 2002) and 336 km3 are associated with international trade in livestock and livestock
products (Chapagain & Hoekstra, 2003a, 2003b, 2003c).
The volumes of water associated with international trade in agricultural commodities
including livestock products are impressive. They are even more impressive when they are
compared with the volumes of freshwater managed in engineered systems for economic use.
Some of the biggest storage structures are in Africa, the Kariba Dam with 180,000 Mm3 stor-
age, the Aswan High Dam with a nominal storage capacity when full of 169,000 Mm3 and a
live storage of about 60% of this volume, and the Akosomba Dam in Ghana with a storage
capacity of 150,000 Mm3 . All these reservoirs have more than four times the capacity of the
Hoover Dam in the USA. Other major structures such as the Three Gorges Dam in China have
a total storage of 80,000 Mm3 and a capacity to manage about 40,000 Mm3 of water partially
to absorb the initial surge of the annual monsoon flood. The three major dams on the Euphrates
in Turkey have a total storage capacity of about 110,000 Mm3 and a live storage of about half
this volume. Most of the 45,000 storage structures (World Commission on Dams, 2000) are
very small compared with these major reservoirs.
If we accept the Rosegrant & Cai (2002) figure of 1430 km3 /yr of water used in irrigation
systems, including the 600 km3 /yr of water pumped from aquifers, mainly for crop production,
then the engineering effort to store water in surface systems affects a relatively limited, though
Virtual water – Part of an invisible synergy that ameliorates water scarcity 143

potentially very productive, proportion of the 5.5 km3 /yr used in agriculture. Soil water is the
major source of water for agriculture and is also the major source of the water associated with
virtual water.

8 THE GLOBAL FRESHWATER RESOURCE, ITS CURRENT AND ESTIMATED


FUTURE AVAILABILITY AND USE

The discussion so far has focused on the economic and trading systems in which water using
activities are embedded to provide solutions to the problems of the water scarce. The history
of water use and management to date across the economies of the world has shown both the
water resources and the systems which remedy the problems of the water scarce have been
sufficient. The other question to answer is whether there will be enough water to meet future
aggregate global water needs using these water managing and trading systems kept in place
by diversifying local economies?
Viewed globally very little of the world’s freshwater is mobilised by human systems. At the
global level a recent study by Rosegrant & Cai (2002: 174), suggests that both the developed
and the developing worlds’ uses of freshwater will only reach 10% of total renewable water
by 2025. It was estimated to be 8% in 1995. Such gross numbers are not safe nor are they of
much relevance for water policy and decision-making.

Table 1. Underlying factors influencing water supply in 1995 and in 2025. Selected countries and
regions.

Storage increase SMAWW increase GMAWW increase Basin efficiency

Countries/ Annual Annual Annual


Regions km3 rate (%) km3 rate (%) km3 rate (%) 1995 2025

Countries
China 157 0.62 145 0.87 27 0.71 0.54 0.68
India 135 1.55 134 0.80 14 0.23 0.57 0.70
USA 0 0.00 48 0.41 3 0.09 0.72 0.78
Regions
South Asia 176 0.90 62 0.66 1 0.05 0.54 0.65
SE Asia 37 0.84 78 1.27 9 1.15 0.47 0.55
Sub-Saharan 74 0.61 50 2.14 20 1.01 0.44 0.50
Africa
Latin America 62 0.47 87 1.15 12 0.61 0.44 0.51
MENA 81 0.72 51 0.66 2 0.12 0.68 0.75
Developing 577 0.66 635 0.88 86 0.46 0.54 0.64
world
Developed 44 0.16 128 0.47 20 0.28 0.64 0.71
world
World 621 0.45 763 0.77 105 0.41 0.56 0.65

Source: Rosegrant & Cai (2002: 172).


SMAWW: surface maximum allowed water withdrawal; GMAWW: groundwater maximum allowed water withdrawal;
MENA: Middle East and North Africa.
144 J.A. Allan

Table 2. Underlying factors/driving forces influencing water demand: projected change 1995 to 2025
for selected countries and regions.

Population increase GDP increase Irrigated area increase Livestock


production
Countries/Regions Million % US$/capita % Million ha % increase %

Countries
China 261 21.3 2390 355 4.9 6.9 122
India 395 42.5 1123 281 9.7 25.1 143
USA 58 21.9 24,405 93 0.7 6.4 31
Regions
South Asia 630 50.7 909 237 11.2 19.1 143
SE Asia 201 42 2332 198 1.7 8.6 136
Sub-Saharan Africa 525 98.4 125 45 3.5 8.6 157
Latin America 212 45.0 3942 110 2.3 25.7 87
MENA 215 66.1 1495 88 1.9 18.4 87
Developing world 2801 47.3 1808 167 26.0 13.5 116
Developed world 57 4.6 17,787 98 2.4 5.1 16
World 2858 37.9 3562 74 28.4 11.9 56

Source: Rosegrant & Cai (2002: 173).

Nevertheless, Rosegrant & Cai (2002) have provided a useful overview first, of driving
forces; and secondly, of surface and groundwater resources in relation to current and projected
sectoral demand (see Table 1 and Table 2). They have made such estimates particularly useful
by incorporating estimates of how increases in water use efficiency will impact future levels of
water use. They include in their models estimated improvements in productive/technical effi-
ciency (Allan, 2002: 129–130) in irrigation, and especially in economic/allocative efficiency
(Allan, 2002: 130).
Table 1 shows that developing economies face different problems from those that affect
developed economies. Developed economies have installed almost all of their surface water
storage structures and will build few such structures in future. Developing economies will sub-
stantially increase the volume of surface water storage by 2025. Surface freshwater sources are
already and will continue to be the main sources of freshwater. Rosegrant & Cai (2002) esti-
mate that surface freshwater will be between six and seven times as important as groundwater
in both developing and developed economies in the coming two decades. Both developing
and developed economies will improve their water use efficiency. Major advances will be
achieved in developing economies because of the much larger volume of freshwater mobilised
in developing economies and the lower base level efficiencies in these economies.
The different demographic driving forces on developing and developed economies are
shown in Table 2. In the North, water demand will not increase significantly in the coming
century because populations have reached a late phase in their demographic transition. The
populations of Southern economies will level off but not until the middle and later decades of
the century (IFPRI, 1995). As a consequence professionals at International Water Management
Institute (IWMI) and the International Food Policy Research Institute (IFPRI) estimate that
there will be significant expansion of irrigated areas in the South and also in the production of
livestock products. The latter are significant as they are very water intensive and are many times
more water intensive in terms of calorie provision than staple crops such as cereals (seeTable 2).
Virtual water – Part of an invisible synergy that ameliorates water scarcity 145

Table 3. Total water withdrawal and total withdrawal as a percentage of total renewable
water (%). Estimated 1995 and projected 2025 for selected countries and regions.

Total withdrawal Total withdrawal as a percen-


contributing to supply tage of renewable water (%)

Countries/Regions 1995 2010 2025 1995 2010 2025

Countries
China 679 771 858 26 30 33
India 674 750 813 30 33 36
USA 497 524 533 24 25 26
Regions
South Asia 1027 1142 1235 24 27 29
SE Asia 203 242 289 4 4 5
Sub-Saharan Africa 128 166 215 2 3 4
Latin America 298 355 411 2 2 3
MENA 236 263 297 69 77 87
Developing world 2762 3145 3528 8 9 10
Developed world 1144 1232 1274 8 9 10
World 3906 4378 4794 8 9 10

Source: Rosegrant & Cai (2002: 174).

Table 3 provides estimates of existing and future levels of withdrawal of freshwater in


developing and developed economies (Rosegrant & Cai, 2002: 174). Those economies that
are known to be most short of freshwater, the economies of the Middle East and North Africa
are already utilising almost 70% of their renewable water. By 2025 they could be using almost
90%. The major economies of Asia, China and India, are using about 30% currently and
by 2025 will be using between 33 and 36%. Sub-Saharan Africa, Southeast Asia and Latin
America have substantial unused water resources. The position for many regions is much worse
than suggested by these data. Most freshwater in natural hydrological systems is inaccessible
to water users. Some significant rivers such as the Nile in Northeast Africa and the Yellow
River in China are subject to such heavy use that they are dry, or almost dry, by the time they
reach the sea. The Nile delivers to the Mediterranean less than 10,000 Mm3 of water out of
the average flow of 84,000 Mm3 . The Yellow River does not reach the sea in most years.
While moderately sized rivers endure very intense use the big rivers deliver to the sea
between 50 and 90% of their natural flows. The Yangtse in China, the Ganges/Brahmaputra
in South Asia and the Congo in Africa have average flows of over 1000 km3 /yr. The massive
Amazon River is in a league of its own draining over 7000 km3 /yr of water to the Atlantic
which is over 95% of its natural flow.
The irrigation sector is the major user of freshwater in any economy where it is decided
to deploy irrigation technologies. It is such a big user that it is possible to generalize, that
wherever it is decided to install irrigation systems the economy will run out of freshwater to
meet the needs of the irrigation sector. As a consequence of the increase in the use of freshwater
in irrigation competition for access occurs at some point unless it is possible to meet the food
needs of a population by food and virtual water imports. The competition is commonly most
evident between water for agriculture and water for the environment if a social movement has
given a voice to environmental water.
146 J.A. Allan

Table 4. Potential and actual consumptive use of water for reliable irrigated agriculture. Estimated
1995 and projected to 2025 for selected countries and regions.

Potential irrigation Actual irrigation Irrigation water supply


consumption consumption reliability index (IWSR)

Countries/Regions 1995 2010 2025 1995 2010 2025 1995 2010 2025

Countries
China 280 283 291 244 227 233 0.87 0.77 0.76
India 400 442 466 321 320 329 0.80 0.72 0.69
USA 133 134 131 124 118 120 0.93 0.88 0.90
Regions
South Asia 605 657 691 484 489 498 0.80 0.73 0.71
SE Asia 98 103 106 85 88 91 0.86 0.82 0.83
Sub-Saharan Africa 69 78 87 50 55 62 0.73 0.65 0.67
Latin America 107 122 129 88 89 96 0.82 0.73 0.75
MENA 156 170 184 122 126 137 0.78 0.72 0.71
Developing world 1445 1549 1617 1162 1165 1214 0.80 0.73 0.71
Developed world 313 314 308 268 263 275 0.86 0.84 0.89
World 1758 1983 1924 1430 1429 1485 0.81 0.75 0.74

Source: Rosegrant & Cai (2002: 177).

As discussed above water scarcity is the result of the allocative decisions made by commu-
nities on the use of land. Water scarcity for the small volumes of water needed for municipal,
industrial and domestic water use is unusual. A few economies in the Arabian Peninsula,
Kuwait, Abu Dhabi, Oman, Qatar, cannot meet such needs but this is exceptional. And
they can easily afford the desalination solution. Cities located at high elevations in the same
region – Damascus (Syria), Amman (Jordan), Sana’a and Ta’iz (Yemen) – also endure current
shortages and face much more serious challenges as sea water desalination is not an economic
option.
Water scarcity to meet the big volumes of water needed to service irrigation schemes
is, however, very common. High water demands for irrigation especially impact the water
needed to sustain the environmental services provided by water, which underpin the security
of all economies. Table 4 shows the potential and actual consumptive use of water for reliable
irrigated agriculture. Again the difference between developing and developed economies is
evident.
The South and the North have different levels of consumption of domestic water and the
economies in the South and the North will increase their levels of use by different proportions.
The use of water and the efficiency of that use in the different livelihoods developed in the
South and the North are very significant with respect to both the volumes of water used
and especially with respect to the capacity of a political economy to achieve economic self-
sufficiency. Once the economic self-sufficiency is achieved, it is possible for communities
and nations to address water scarcity.
The discussion of water availability and water use and the different trajectories of use
and development evident in the South and the North have a number of recurring themes.
These recurring issues are the consequence of the different levels of population change and
of socio-economic development in the South and the North. The South has five sixths of
Virtual water – Part of an invisible synergy that ameliorates water scarcity 147

the world’s population. The North has only about one sixth. The South has limited adaptive
capacity reflected in average GDPs per head (see World Bank: www.worldbank.org) and
Human Development Indices (HDI, see UNDP: www.undp.org) which are only 10% of those
enjoyed in the North. Adapting to water scarcity has proven to be possible in the diverse and
strong economies of the North whether they enjoy secure or insecure water. The economies of
the South are poor and it is this poverty which impairs their capacity to adapt to water scarcity.
The demographic and resource numbers and the ability to estimate the pace and the effect-
iveness of socio-economic development mean that analysts of water futures face uncertainty
and therefore ambiguity. Pessimists emphasise the uncertainty that the global water resource
will be sufficient to meet the water demands of the global population. Optimists observe the
declining trends in population increase and the relatively modest challenge of increasing food
production by 30% or at worst 50%. Yield increases achieved in the second half of the 20th
century if extended to the poorly performing farms of sub-Saharan Africa would easily meet
future food needs.

9 VIRTUAL WATER AND THE THREEFOLD SYNERGY ATTENUATE THE


MYTH OF THE WATER CRISIS, BUT PROVIDE SIMULTANEOUSLY ONE OF
THE MAJOR SOLUTIONS

The purpose of the chapter has been to highlight some important underlying fundamentals in
the domain of water use and water policy-making. Three ignored and de-emphasised elements
of the hydrological and the hydro-economic system have been identified. This threefold syn-
ergy – soil water, virtual water and the socio-economic development – enables economies to
trade their way to water security. All three elements are ignored by water policy-makers and
especially by those who promote the idea that there is severe and possibly inaddressable global
water crisis. First, soil water is ignored as part of water budgets. Secondly, virtual water is de-
emphasised because if it were to become prominent in the political discourse of water scarce
economies it would bring negative and destabilising political consequences. Those affected
would be the major water users in the irrigation sector and the water policy-making elite in the
poor economies that cannot enjoy the policy-making benefits deriving from economic diver-
sification. Thirdly, economic diversification which is not intuitively seen as directly relevant
to the water sector as it plays a hidden-hand role in enabling virtual water transactions. The
technical professionals who generally manage water do not see solutions outside the water
sector. As a result they are very exercised about the inadequacies of their local watersheds.
The politically rational de-emphasis and lack awareness of these underlying fundamentals
is sub-optimising for the water using and water policy-making processes. They aggravate the
already existing, seriously second best, water managing operations, especially in developing
economies. They also enable water policy makers to avoid necessary, but politically costly, re-
allocative reforms. Such reforms would both improve the economic returns to water as well as
enhance the levels of environmental service provided by water in the natural water environment.
The threefold synergy in the global political economy of water has not been highlighted
until the writing of this chapter. The author senses that the identification of three economically
invisible and politically silent processes in the water sector has already been identified as
normal processes in highly diversified and developed political economies. Certainly it is not
difficult to share these ideas with water users from developed economies. In economies where
148 J.A. Allan

poverty prevails, the ideas have no purchase because the power of economic diversification
has not been experienced.
Myths establish themselves in circumstances of uncertainty. They flourish and endure as
they have been constructed to accord with cultural preferences. Contradictory myths com-
monly exist at the same time. The myth of future unmanageable global scarcity has been
generated in the risk aware neo-liberal North in late-modernity – that is since the late 1970s.
The global water crisis myth is proclaimed at the same time that the farmers and politicians of
Middle Eastern economies claim that all they need is a little more water, then they will manage
it more carefully, and everything will be all right. It has been argued here that both these myths
are wrong. They have been constructed in the first case by scientists with a partial understand-
ing of the complex hydro-economic global system. They emphasise demographic indicators
and water resource availability and ignore the remedies outside the hydrological systems. In
the second case where farmers and their political leaders face palpable water shortages, in for
example the Middle East, there is a convergence on the myth that water resources are adequate
if differently managed.
Myths usually endure because people and politicians want to ignore bad news and avoid
change. This chapter has highlighted that such myths are kept in place in water scarce countries
because there are invisible economic processes that make it appear that water is not a problem.
The myth of global water scarcity is a different type of myth. It is the result of hyper-risk aware
Northern environmentalists and water scientists relating local hydrological and demographic
data without taking into account remedies available in technology, economic development and
trade. The purpose of the chapter has been to show that non-hydrological global systems have
been ameliorating local water scarcity for half a century. They have the capacity to meet the
water demands of foreseeable future populations.

REFERENCES

Allan, J.A. (2001). The Middle East water question: hydro-politics and the global economy. London:
I B Tauris.
Allan, J.A. (2002). Hydro-peace in the Middle East: Why no water wars? A case study of the Jordan
River Basin. SAIS Review, Vol. XXII: 2 (Summer–Fall, 2002): 255–272.
Allan, J.A. (2003). Virtual water: the water, food, and trade nexus. Useful concept or misleading
metaphor? Water International, 28(1): 4–11.
Allan, J.A. & Olmsted, J. (2003). Trading virtual water, the MENA water deficit and the consequences
for water policy reform. In: Lofgren, H. (ed.), The Middle East water economy. IFPRI, Washington.
Bowen, H.P.; Leamer, E.E. & Sveikauskas, L. (1987). Multicountry, multifactor tests of the factor
abundance theorem. The American Economic Review, 77(5), December: 791–809.
Brichieri-Colombi, J.S.A. (2003). Food security, irrigation and water stress: logical chain or environ-
mental myth? Occasional Paper 57, King’s College London/SOAS Water Research Group. London,
UK. (See http://www.kcl.ac.uk/kis/schools/hums/geog/water/occasionalpapers/home.html#trade).
Brichieri-Colombi, S. (2004). Who speaks for the river? Unpublished PhD Thesis of the University of
London.
Brown, L. (2003). Plan B: rescuing a planet under stress and a civilization in trouble. WW Norton,
New York, USA.
Chapagain, A.K. & Hoekstra, A.Y. (2003a). Virtual water between nations in relation to trade in livestock
and livestock products. Value of Water Research Report Series, 13. IHE, Delft, The Netherlands.
Chapagain, A.K. & Hoekstra, A.Y. (2003b). The water needed to have the Dutch drink coffee. Value of
Water Research Report Series, 14. IHE, Delft, The Netherlands.
Virtual water – Part of an invisible synergy that ameliorates water scarcity 149

Chapagain, A.K. & Hoekstra, A.Y. (2003c). The water needed to have the Dutch drink tea. Value of Water
Research Report Series, 15. IHE, Delft, The Netherlands.
Dyson, T. (1996). Population and food: global trends and future prospects. Routledge, London, UK.
Dyson, T. (2001). World food trends: a neo-Malthusian prospect? Proceedings of the American
Philosophical Society, 145(4), December: 438–455.
Earle, A. (2001). The role of virtual water in food security in Southern Africa. Occasional Paper 33.
King’s College London/SOAS Water Research Group, London, UK. (See http://www.kcl.ac.uk/kis/
schools/hums/geog/water/occasionalpapers/home.html#trade).
El-Fadel, M. & Maroun, R. (2003). The concept of virtual water and its applicability in Lebanon. In:
A.Y. Hoekstra, Virtual water trade: a quantification of virtual water flows between nations in relation
to international crop trade. Value of Water Research Report Series, 12. IHE, Delft, The Netherlands:
171–182.
FAO (2003). Review of world water resources by country. Water Report, 23. FAO, Rome, Italy.
FAO (2004). Aquastat database. FAO, Rome, Italy (On-going on: http://www.fao.org).
Fine, B. & Saad-Filho, A. (2003). Marx’s Capital. Fourth edition. Pluto Press, London, UK.
Goodchild, P. (2004). Meet the real Dr. Strangelove. The Guardian, G2 Supplement, 2 April 2004: 8–9.
Haddadin, M.J. (2003). Exogenous water: a conduit to globalization of water resources. In: A.Y. Hoekstra,
Virtual water trade: a quantification of virtual water flows between nations in relation to international
crop trade. Value of Water Research Report Series, 12. IHE, Delft, The Netherlands: 159–169.
Hakimian, H. (2003). Water scarcity and food imports: an empirical investigation of the virtual water
hypothesis in the MENA Region. Occasional Paper 46. King’s College London/SOAS Water Research
Group. London, UK. See (http://www.kcl.ac.uk/kis/schools/hums/geog/water/occasionalpapers/home.
html#trade).
Harkness, J.P. (1978). Factor abundance and comparative advantage, The American Economic Review,
68, December: 784–800.
Hoekstra, A.Y. (ed.) (2003). Virtual water trade: Proceedings of the international expert meeting on
virtual water trade. Value of Water Research Report Series, 11. IHE, Delft, The Netherlands.
Hoekstra, A.Y. & Hung, P.Q. (2002). Virtual water trade: a quantification of virtual water flows between
nations in relation to international crop trade. Value of Water Research Report Series, 11. IHE, Delft,
The Netherlands.
IFPRI (1995). Global food security. IFPRI Report, October 1995. International Food Policy Research
Institute, Washington, D.C. See (www.cgiar.org/ifpri).
IFPRI (1997). China will remain a grain importer. IFPRI Report, February 1997. International Food
Policy Research Institute, Washington, D.C. See (www.cgiar.org/ifpri).
Krueger, A.O. (1977). Liberalization attempts and consequences. Ballinger Publishing Co., Cambridge,
Massachusetts, USA.
Krugman, P.R. & Obstfeld, M. (1991). International economics – theory and policy. HarperCollins
Publishers, New York, USA.
Lawler, K. & Seddighi, H. (2001). International economics – theories, themes and debates. Prentice
Hall, London and New York.
Merrett, S. (2003). Virtual water and Occam’s razor. Water International, 28(1): 1–3, 12–16.
Postel, S. (1999). Pillar of sand. Norton, New York, USA.
Postel, S. & Richter, B. (2003). Rivers for life: managing water for people and nature. Island Press,
St. Louis, Washington University, USA.
Renault, D. (2003). Virtual water: principles and virtues. In: A.Y. Hoekstra, Virtual water trade: a
quantification of virtual water flows between nations in relation to international crop trade. Value of
Water Research Report Series, 12. IHE, Delft, The Netherlands: 77–91.
Rosegrant, M.; Agcaoli, M. & Perez, N.D. (1995). Global food projections. Diane Publishing Company,
Washington, USA.
Rosegrant, M. & Cai, X. (2002). Global water demand and supply projections. Part 2: results and
prospects to 2025. Water International, 27(2): 170–182.
Shiklomanov, I.A. (2000). Appraisal and assessment of world water resources. Water International,
25(1): 11–32.
Stauffer, T. (2004). The economic limits of seawater desalination. Unpublished draft.
150 J.A. Allan

Wakemagel, M. & Rees, W. (1996). Our ecological footprint: reducing human impact on the Earth. New
Society Publishers, Gabriola Island, BC, Canada.
Weber, M. (1904 and 1917). The methodology of the social sciences. Translated and edited by Edward A.
Shils and Henry A. Finch. Published in 1949. Free Press, New York, USA.
Wichelns, D. (2004). The policy relevance of virtual water can be enhanced by considering comparative
advantages. Agricultural Water Management, 66: 49–63.
World Bank (2003). World Bank water resources sector strategy. World Bank, Washington, USA. See
World Bank website.
World Bank-WRI (2003). World Resources 2002–2004, decisions for the Earth: balance, voice, and
power. World Bank and World Resources Institute. United Nations Development Programme, United
Nations Environment Programme, New York and Washington, USA.
World Commission on Dams (2000). Dams and development. Earthscan Publications, London, UK.
Zimmer, D. & Renault, D. (2003). Virtual water in food production and global trade: review of methodo-
logical issues and preliminary results. In: A.Y. Hoekstra, Virtual water trade: a quantification of virtual
water flows between nations in relation to international crop trade. Value of Water Research Report
Series, 12. IHE, Delft, The Netherlands: 93–109.
CHAPTER 9

Virtual water – Part of an invisible synergy that ameliorates


water scarcity: commentary

J. Ramirez-Vallejo
Harvard University, Cambridge, Massachussets, USA

ABSTRACT: Prof. Allan (this volume) argues that there is sufficient water in the world to meet
current and future water needs, mainly because there are three processes in what he calls a threefold
synergy: soil water, virtual water and the socioeconomic development that enables economies to trade
their way to water security. In these comments, I argue that the virtual water argument does not hold
true and that, although it is a useful concept to remind us of the water scarcity in some regions of the
world, it is a concept that does not have significant implications on water and agricultural policy. These
comments also include some reflections about the question regarding a possible water crisis in the future,
the supply and demand of water resources, and the EU/USA agricultural subsidies. Finally, I present a
hypothesis as why water resources development has not reach the top of the international agenda.

Keywords: Virtual water, demand, agricultural subsidies, soil water

1 INTRODUCTION

In his chapter, Prof. Allan (this volume) argues that there is sufficient water in the world to
meet current and future water needs, mainly because there are three processes in what he calls
a threefold synergy –soil water, virtual water and the socioeconomic development that enables
economies to trade their way to water security. Therefore, according to Allan, the expected
water crisis would eventually be solved. Then he develops each of these processes, some in
more depth than others, and highlights several underlying fundamentals in the domain of water
use and water policymaking.
The three processes are: (1) soil water, which according to him, is ignored by water budgets
around the world; (2) virtual water, or water that is embedded in the international trade of
agricultural products, mainly in meat and cereals; and (3) economic diversification or socio-
economic development, which acts as a hidden-hand in enabling virtual water transactions.
With these new processes in mind, Professor Allan, the father of the virtual water concept,
presents an optimistic scenario of the future world water balance (Allan, 1996).
My comments are presented in the following order. First, I argue that the virtual water
argument does not hold true and that, although it is a useful concept to remind us of the water
scarcity in some regions of the world and the amount of water used in agriculture, it is a concept
that in a world in which water subsidies and tariff and non-tariff measures will continue to
be the norm, does not have significant implications on water and agricultural policy. In the
second part of my comments, I will discuss the major question regarding a possible water
crisis in the future. In particular, I try to explore the answer from the supply and demand side
152 J. Ramirez-Vallejo

of water resources. Third, I will comment on the soil-water and socioeconomic development
ameliorating processes identified and proposed by Prof. Allan, and will make a note on the
EU/USA agricultural subsidies as a potential fourth process suggested by the author. Finally,
I will present a hypothesis as why water resources development has not reach the top of the
international agenda.

2 THE VIRTUAL WATER ARGUMENT

The virtual water concept comes from the idea that water should be treated as a production
factor and is equivalent to the volume of water needed to produce a commodity or service. The
virtual water argument, on the other hand, shows that the importation of agricultural products
that require important amounts of water represents the importation of water into a water scarce
country. Using Prof. Allan’s words (this volume), “trade in commodities has already achieved
a pivotal position in enabling the comparative advantages and disadvantages of regional water
endowments to be balanced”.
Under the virtual water argument, food trade then becomes an instrument to augment water
supplies on the scale needed to meet the domestic food demand, which in trade-theory terms,
is equivalent to saying that the Heckscher-Ohlin (H-O) theorem holds when applied to water
as a key input in the agricultural production process.
Leontief (1953) first employed the factor content of trade approach in his well-known
test of the H-O theorem, which states that a country exports products that use intensively a
relatively abundant input. However, the proof of this theorem implies several assumptions
such as factors can move without cost among industries within a country, but are completely
immobile internationally, which might be the case of water as a production factor. Other more
restricted assumptions are that the production functions for all countries (products) exhibit
constant return to scale and each country has the same productive technology for each good,
and consumers have the same utility functions, which clearly is not the case for the use of
water in agriculture worldwide.
Does the H-O theorem apply to water-dependent agricultural trade? To answer that question
Ramirez-Vallejo & Rogers (2004b) ran a test and found that the proposition of conformity in
sign between water contents and excess water shares receives relatively little support. Using
Fisher’s Exact Test, the hypothesis of independence between sign of the water contents and
of the excess water shares could not be rejected at the 95% level. They also tried to correct
for differences in productivity of land and found similar results. This should not come as a
surprise because this type of result has become the norm when testing the H-O hypothesis,
Deardorff (1982), Ethier (1984) and Helpman (1984).
This finding again proves one more time, what Bhagwati (1964) states in his survey article:
“Where the theory [H-O theorem] is deficient is in its not having been extended to a multi-
country framework. The prediction holds when there are many commodities and two countries.
However, when there are many countries and commodities it is not clear, in the absence of theo-
retical analysis, what conditions will suffice to make the Heckscher-Ohlin hypothesis true…”.
The main reason for the failure of the H-O theorem is that under conditions sufficient
to guarantee factor-price equalization there exist many efficient production configurations
in the world consistent with the equilibrium factor prices and a given distribution of factor
endowments among countries.
Virtual water – Part of an invisible synergy that ameliorates water scarcity: commentary 153

On the inability of the virtual water argument, Wichelns (2004) recognizes that the virtual
water metaphor addresses resource endowments, but does not address production technologies
or opportunity costs. More specifically, the price equalization hypothesis that underlies the
H-O theorem rarely applies in the case of water as an agricultural input. Values of water differ
significantly from one country to the next, and even within countries. The price for water that
is actually paid by farmers and that is internalized in the farmer’s decision process is distant
from the true opportunity cost of water.
Under these circumstances, the farmer’s choice of whether to plant water-demanding crops
does not follow the water scarcity signal of the need to choose less-water-demanding crops.
For instance, in the irrigation districts in Northern Mexico, farmers continue to cultivate cer-
eals and water demanding crops (high levels of virtual water) because of the highly subsidized
crop prices and low water fees administered by the government. In a scenario of no agricul-
tural subsidies, the farmers in these districts should have switched from cereals to fruits and
vegetables or to higher-value crops after NAFTA became effective to take advantage of the
opportunities presented by the USA market.
If water endowment does not explain the direction of agricultural trade or virtual water
trade, what does? The truth of the matter is that complex interactions among technology, policy,
investment, environment, and human behavior influence the trade of virtual water. Indeed,
any factor or condition that alters the demand or supply sides of the agricultural commodities
markets impact agricultural trade pattern and therefore on the Net Virtual Water Trade. Among
these factors are: trade agreements; economic and population growth; technological innov-
ation; subsidies to agricultural products and inputs; international prices of agricultural products
and inputs; macroeconomic policies in import and export countries; domestic microeconomic
policies; political and economic crisis; degradation of natural resources; and cultural and
religious changes and constraints (Rosegrant & Paisner, 2001; Ramirez-Vallejo & Rogers,
2004a).
On the other hand, the embodied virtual water in agricultural trade flows are explained by
the variables considered by economic agents at the time production and trade decisions are
made. These decisions are linked directly to price, product characteristics and real demand
conditions. In other words, the decision taken by economic agents when trading agricultural
products is directly related to attributes of the agricultural products, including their price,
rather than to the water used for production alone. Rather water enters in the decision process
as a major factor of production, but only indirectly.
Ramirez-Vallejo & Rogers (2004a) formally researched the determinants behind agricul-
tural imports in order to understand the hidden factors that indirectly drive virtual water flows.
They found, for example, a very strong relationship between income and virtual water imports
with an income elasticity of 0.52 for the year 2000–2001. This means that in a scenario of
a world two times richer than today, the world could experience an increase in virtual water
flows of 50%, or close to 400 km3 . However, it is difficult to tell if this virtual flow would
occur in the direction from water surplus countries to water deficit countries.
Income has a double role in virtual water. First, higher income and purchasing power holds
out the possibility of being able to offset most of the constraints that prevent access to water in a
particular place at a specific time. In addition, higher national income increases the possibility
of importing food from other countries.
Ramirez-Vallejo & Rogers (2004a) also found that the level of support or protection to
agriculture is also significant in explaining virtual water demand. Elasticity for the support
154 J. Ramirez-Vallejo

100

90

80

70 Other
Percentage of Trade

60
Emerging
50

40
Transition
30

20
OECD
10

0
Imports 1993 Imports 1998 Exports 1993 Exports 1998

Figure 1. Direction of Agricultural Trade before and after the Uruguay Round.

of agriculture was found to be −0.9. The higher the support for agriculture in one country
compared to other countries, the less the amount of virtual water imports1 .
If one examines at who trades agricultural goods with whom, one finds that OECD coun-
tries trade mostly with each other. In 1998, 83% of OECD agricultural imports came from
other OECD countries, while 85% of agricultural exports went to other OECD countries (Fig-
ure 1). There has been little penetration of OECD markets by less developed countries, whose
combined share of OECD agricultural imports increased from 8% to 9% during the period of
most significant agricultural liberalization around the world.
On the other hand, developing countries are sending an increasing share of their agricultural
exports to other developing countries. One reason for this is that, on average, developing coun-
tries are growing more rapidly than developed ones. At the same time, the increasing share of
non-bulk food trade in total food trade reinforces the reorientation of trade, since protection in
developed countries is often higher for semi-processed and processed products (OECD, 2001).
In the case of emerging economies, they show a more diverse profile of trading relationships.
It is interesting to note that Brazil sends half its exports to the European Union (EU), compared
with just 27% in the case of Argentina. Nearly two-thirds of South African exports go to OECD
countries, not least because of the underdevelopment of regional African markets.
The concept of virtual water is then appealing to educate public officials and society in
general that water in some parts of the world is a scarce resource and that agriculture uses
the great majority of water resources available on Earth. The argument also has an implicit
lesson underscoring the importance of running efficiently irrigation districts so water could be
allocated to other uses including ones benefiting the environment. However, the virtual water

1
Other variables were found to have some explanatory power of the variance of virtual water imports.
These are: (1) the average income (GDP); (2) population; (3) agriculture as value added (% of GDP); (4)
irrigation (actual and potential); and (5) exports of goods and services (% of GDP). The sign found for
each of these variables in the estimated equation were as expected, with the absolute value of elasticities
ranging from 0.3 to 0.7.
Virtual water – Part of an invisible synergy that ameliorates water scarcity: commentary 155

argument, if applied improperly, could send the wrong message in terms of policy making in
agriculture and water resources. For instance, a country could delay important investments
now and decide instead to import foodgrains, or, it could choose not to remove price subsidies
with the objective of saving water.

3 IS IT A DEMAND OR A SUPPLY PROBLEM?

3.1 Supply or demand?


As Professor Allan states, “optimists and pessimists interpret the underlying fundamentals
differently”. I consider myself a demand-driven pessimistic and a supply-driven optimistic
unless some unlikely structural changes occur in policy.
Professor Allan argues in his chapter that there is sufficient freshwater and soil water in
the world to meet current and future water needs if we accept the demographic predictions. I
agree with Prof. Allan that the world does have the water resources to meet the needs of future
populations, if one considers today’s technology to mobilize it and to treat it in such a way
that it will become available where it is needed.
Agriculture accounts for more than 70% of total water use in most countries. Therefore, the
issue whether there is enough water is related to the question of whether there is enough water
to feed the population and to end hunger and prevent recurrence of famine and starvation.
IFPRI estimates that roughly 800 million people are chronically undernourished today –three
times the population of the USA– and unfortunately nothing tells us that there will be a change
in the trend in the near future (IFPRI, 2003).
The problem is not water supply when one could most probably induce the generation
and use of sophisticated technologies to obtain enough water to satisfy water needs. With
increasing pressure on use of natural freshwater in parts of the world, other non-conventional
sources of water are growing in importance: (1) the production of freshwater by desalination
of brackish or saltwater (mostly for domestic purposes); and (2) the reuse of urban or industrial
wastewaters (with or without treatment), which increases the overall efficiency of use of water
(extracted from primary sources), mostly in agriculture, but increasingly in the industrial and
domestic sectors. This category also includes agricultural drainage water.
Although the unit cost of these technologies has been decreasing over time, they remain too
expensive for most communities in the developing world. Thus, the real problem comes from
the demand side as to whether there would be enough economic resources at the individual,
regional or country levels to pay for the water/technology needed. The resources will be needed
not only to cover the cost of water treatment but also its transportation to users. Table 1 shows
how significant is the per capita annual water withdrawn by income groups for different uses
in Africa. The water withdrawn per capita for the higher income group is on average three
times larger than the amount of the lower income group.

3.2 Induced innovation hypothesis


If demand increases and local water supplies were scarce, then this situation would most likely
induce a technology change. Hayami & Ruttan (1985) have presented various tests to their
well-known induced innovation hypothesis in which they find that technical and institutional
changes are induced through the responses of farmers, agribusiness entrepreneurs, scientists,
156 J. Ramirez-Vallejo

Table 1. Sectoral water withdrawals by income group in Africa.

Per capita annual withdrawal (m3 )


Country
income group Total Agriculture Domestic Industry

Low income 385 351 (91%) 15 (4%) 19 (5%)


Middle income 454 313 (69%) 59 (13%) 82 (18%)
High income 1167 455 (39%) 163 (14%) 549 (47%)

Source: World Bank (www.worldbank.org).

and public administrators to resource endowments and to changes in the supply and demand
of factors and products.
Using their findings, one could say that the state of relative endowments and accumulation
of the primary resources of land, labor, and water, would be a critical element in determining
the pattern of technical change that will occur in water resources and agriculture in the future.
In other words, technical change embodied in new and more productive inputs may be induced
primarily either to save labor, to save land or to save water. If water were priced at its real
opportunity cost, then technology changes would likely generate a lower-cost access to water
resources. However, if water continues to be undervalued by society or highly subsidized then
future technological changes will occur in other directions.

3.3 Water and food security


I agree with Prof. Allan that whether water will become a major constraint to the achievement
of food security in many developing countries remains to be seen. However, I find it difficult
to believe that by moving water resources virtually some countries would be able to solve their
food problems regionally or at the national level. In his classic book Transforming Traditional
Agriculture, T.W. Shultz (1964) suggested that significant growth in productivity cannot be
brought about by the reallocation of resources in traditional agricultural systems. Significant
opportunities for growth will become available only through changes in technology –new
husbandry techniques, better seed varieties, more efficient sources of power, cheaper and
more effective nutrients, and biotechnology.
It is difficult to disagree with Prof. Allan that economic diversification tends to reduce
relative water scarcity. The economies of the globe usually adjust from the demand side and
generate the required income for the population to effectively demand higher priced water. The
example given by Prof. Allan in which job creation outside agriculture has generated incomes
and revenues with which economies have been able to access virtual water, is a classic example
of this type of adjustment of water demand through economic diversification.

4 COMMENTS ON THE OTHER TWO INVISIBLE PROCESSES: SOIL


WATER AND SOCIOECONOMIC DEVELOPMENT

Prof. Allan introduces for the first time two processes that, in his opinion, have made possible
for many communities around the world to access water without having to go to war. These
processes substitute water in space and time when water is needed. One is the soil water
Virtual water – Part of an invisible synergy that ameliorates water scarcity: commentary 157

about which he argues has not been taken into account when observing water supply at the
country level. While the second process is the socioeconomic development process where
diversification experienced by some economies has increased income and made possible
buying intensive-water-agricultural products from other areas and therefore saving local water
resources.

4.1 Soil water


Prof. Allan argues that soil water is not taken into account when computing national water
endowments by international agencies. However, FAO, in its AQUASTAT information system,
begins estimating the total amount of water from rainfall and then, after applying several
assumptions, estimates the amount of surface and groundwater, external water resources and
other uses. FAO defines available water resources as “water net balance in a given state of use
and exploitation of the resources and not with a meaning of water offer. The availability may
be: (1) equal to resources minus withdrawal at the local level of a subsystem, where a part of
the water withdrawn cannot be returned into the system; or (2) equal to resources minus final
consumption at a more regional scale (watershed, country), where the balance encompasses
all the use systems” (FAO, 2002). If this methodology and these definitions are applied, then
soilwater has always been considered as part of the water endowment or water availability of
a country or region. This is seen even more clearly when looking at the water balance usually
computed in hydrologic models used in watersheds.
For the sake of clarification, plants take only one part of soil water as a source of growth,
the part denominated as capillary water. The other components are gravitational water and
hygroscopic water which are not used by plants for a variety of reasons. Gravitational water
drains quickly from the soil, and hygroscopic water remains adhered to the soil particles,
therefore could be considered constant over time. The available water to the plant roots is the
one located between the wilting point and the field capacity, and as explained above is part of
the water balance used to estimate water endowments at the country and regional levels.

4.2 The input and output substitution effects: a forth process?


The input and output substitution effects is a process that could save water by obtaining
more food per drop, or producing less water demanding products for the new preferences
of consumers. In the case of inputs, the water required to produce food would be less if
it were feasible to use other inputs that in some quantity may reduce water requirement. For
example, a higher amount of fertilizer applied might increase yield with the same or less water
requirements. In this case, fertilizer generates the concept of opportunity cost of virtual water.
The same could be said for other inputs such as seeds, land, labor, machinery, and even financial
resources. This is why if the prices are right, the correct signal is given to the innovation process
to obtain high cost input saving technologies. In the case of output, a change in the daily intake
diet preferences could change significantly in terms of water requirements. For instance, a
switch from meat to fruits and vegetables would save significant amounts of water.

4.3 EU/USA agricultural subsidies: another ameliorating forth (the forth force?)
Prof. Allan points out the agricultural policies of the EU and the USA as a process that has led
to the water scarcity since the middle of the 20th century. Based on his calculations, around
158 J. Ramirez-Vallejo

25% of total virtual water trade is a result of the lower world prices for grain produced by
subsidies. This is true; however, the real effect could be overestimated because the increase
in world prices of commodities in a world without subsidies could be significant2 (Valdez &
Zietz, 1995; OECD, 2000).
However, the level of protection to agricultural products via tariffs and duties and non-
tariff instruments by all countries, developed and developing, has distorted the virtual water
movement worldwide. Ramirez-Vallejo & Rogers (2004a) showed that, for example, using
IFPRI’s IMPACT model results, a scenario of full liberalization of agriculture compared to a
baseline scenario would have a greater net effect of virtual water flows from the relocation
of meat trade than from the adjustment in cereals’ trade. When the net effect of the meat
and cereals markets are added together, the two major contributors to the increase in virtual
water trade would be the USA, which would increase its annual virtual water exports in about
86 km3 , and Latin America would have a similar increase: 89 km3 . These are the two water
surplus regions in the world. The major changes in virtual water imports would occur in Asia
in general (South Asia, Southeast Asia, East Asia) with an increase of 112 km3 , Sub-Saharan
Africa with an increase of almost 40 km3 and the former Soviet Union with an increase in
water imports of 22 km3 , mostly because of an increase in meat imports. West Asia and North
Africa together, on the other hand, would decrease the level of virtual water imports to about
7 km3 , but would remain as an important net importer of virtual water of about 176 km3 .

5 A FINAL HYPOTHESIS

Finally, I end my comments with a hypothesis on why water has not reached the top of the
international agenda. Besides the lack of consensus on what the message on water develop-
ment policy should be among water scientists3 , the challenge ahead is how to bring about
significant shifts in perception of the role of water in poverty alleviation and competitiveness
enhancement. As long as water is not considered important in alleviating poverty and increas-
ing competitiveness, it will be very difficult to ask for something more from what Prof. Allan
terms the political economy superstructure of water related decision making.

REFERENCES

Allan, J.A. (1996). Water Use and Development in Arid Regions: Environment, Economic Development
and Water Resource Politics and Policy. Water Use and Development, 5(2): 107–115.
Bhagwati, J. (1964). The Pure Theory of International Trade: A Survey. Economic Journal, LXXIV:
1–84.
Deardorff, A.V. (1982). The general validity of the Heckscher-Ohlin Theorem. American Economic
Review, 72(4): 683–694.

2
IFPRI found that full liberalization would have a significant effect on cereal prices by 2020 with rice
increasing the most at 14%, followed closely by maize, wheat and other course grains. Meat prices would
respond with even sharper price increases with beef alone being subject to an 18% increase.
3
Some water specialists still argue in favor of more water subsidies in order to increase the supply of
water resources infrastructure to the lower income segment of the population, while others, on the other
hand, argue that water should be priced at the opportunity cost of the resource no matter how high this
might be.
Virtual water – Part of an invisible synergy that ameliorates water scarcity: commentary 159

Ethier, W.J. (1984). Higher dimensional issues in trade theory. In: R.W. Jones & P.B. Kenen (eds.),
Handbook of International Economics, Vol. I. Amsterdam, the Netherlands: 131–184.
FAO (2002). Review of Water Resources by Country, Water Report 23. Food and Agricultural
Organization, United Nations.
Hayami, Y. & Ruttan, V. (1985). Agricultural Development: An International Perspective. Revised and
Extended Edition. Jhon Hopkins University Press. Baltimore, Maryland, USA.
Helpman, E. (1984). The factor content of foreign trade. Economic Journal, 94: 84–94.
IFPRI (2003). Annual Report 2002–2003, Trade Policies and Food Security. International Food Policy
Research Institute, Washington, D.C., USA.
Leontief, W.W. (1953). Domestic Production and Foreign Trade: The American Capital Reposition Re-
examined. Reprinted in: R.E. Caves & H.G. Johnson (eds.), Readings in International Economics
(London, Allen & Unwin, 1968).
OECD (2000). Agricultural Policies in Emerging and Transition Economies.
OECD (2001). Agricultural Policies in OECD Countries: Monitoring and Evaluation.
Ramirez-Vallejo, J. & Rogers, P. (2004a). Virtual Water Flows and Trade Liberalization. Journal of Water
Science and Technology, 49(7): 25–32.
Ramirez-Vallejo, J. & Rogers, P. (2004b). Mexico: NAFTA, Virtual Water, and the Economic Value of
Water (mimeo). DEAS, Harvard University, USA.
Rosegrant, M.W. & Paisner, M.S. (2001). Global Food Projections to 2020. International Food Policy
Research Institute, Washington, D.C., USA.
Shultz, T.W. (1964). Transforming Traditional Agriculture. Yale University Press. New Haven,
Connecticut, USA.
Valdez, A. & Zietz, J. (1995). Distortions in World Food Markets in the Wake of GATT: Evidence and
Policy Implications. World Development, 23(6): 913–926.
Wichelns, D. (2004). The policy relevance of virtual water can be enhanced by considering comparative
advantages. Agricultural Water Management, 66: 49–63.
V

Groundwater
CHAPTER 10

Significance of the Silent Revolution of intensive


groundwater use in world water policy

M.R. Llamas
Royal Academy of Sciences, Madrid, Spain

P. Martínez-Santos
Complutense University of Madrid, Spain

ABSTRACT: A series of fairly new factors such as virtual water, the rise of desalination technologies
and intensive groundwater use currently look as though they will exert a strong influence on future water
policy. This chapter is concerned with the latter of the three: the Silent Revolution of intensive groundwater
use in arid and semi-arid countries. Over the last half century, millions of farmers have independently
drilled their own wells in the pursuit of the socio-economic advantages of groundwater irrigation. This
has been due to fairly recent advances in well drilling and pumping, which together with the development
of hydrogeology as a solid body of science, have made groundwater more widely available. The intrinsic
benefits of groundwater irrigation in relation to traditional surface water systems, such as the ready
availability of the resource or the resilience of aquifers against drought, constitute the main reason behind
the spectacular groundwater development of many arid and semi-arid countries worldwide. Despite these
undeniable benefits, certain problems (mainly related to groundwater quality degradation and water table
depletion) have also arisen in some places. While no two cases are the same, a pattern of events can be
observed in many of these regions, thus leading to the conceptualisation of this intensive groundwater-
based development. Thus, five stages can be distinguished: Hydroschizophrenia, changes in water policy
due to the Silent Revolution, Farmer Lobbies, Conservation Lobbies and Social Conflict. In any case,
despite the significant role groundwater development is already playing in the eradication of poverty as
well as towards fulfilling the United Nation’s Millennium Goals, it cannot be seen as a panacea to solve
all the world’s water problems. These need to be dealt with on a case-by-case basis, in order to achieve an
adequate conjunctive management of surface and groundwater resources. Finally, there is a real need to
assess and correct the traditional imbalance in favour of conventional surface water systems that exists
in most water agencies, and which is the main cause behind important social conflicts.

Keywords: Intensive groundwater use, silent revolution, irrigation efficiency, UN Millennium Goals

1 INTRODUCTION

Currently, three issues pose an open challenge to the widely accepted paradigms of world water
policy: the concept of virtual water, the improvements in desalination technology and the Silent
Revolution of intensive groundwater use for irrigation. Abundant references to the first two
already exist in scientific literature and have been widely discussed during the Santander
Workshop (14–16 June 2004). This chapter is only concerned with the global implications of
the latter.
164 M.R. Llamas & P. Martínez-Santos

In the last few decades, millions of farmers in arid and semi-arid countries have, at their
own expense and risk, drilled their own wells in order to make use of the intrinsic advantages of
groundwater-based irrigation. This phenomenon, the Silent Revolution, has thus been carried
out with little or no planning on the part of governmental water agencies. The benefits of
this Silent Revolution have been quite significant, even if some groundwater-related problems
have also arisen in certain places. Precisely these have caused certain scholars and water
policy makers to spread a series of hydromyths, which have led some to consider groundwater
development a pillar of sand prone to collapse, or as a bubble, likely to burst (Postel, 1999).
However, while these problems are sometimes real, it can be shown that they are many times
pretended or exaggerated, or due to poor land use planning, rather than to intensive groundwater
development.
The fact is that intensive groundwater use has been contributing for years to achieve two of
the main goals set by the United Nations (2000) for the new Millennium: to halve by 2015 the
number of people worldwide who suffer from malnutrition or do not have affordable access
to drinking water. It must also be noted that fulfilling the another goal of the Millennium
Declaration, i.e. to cut down by 50% the number of people living under the poverty threshold
(US$ 1–2 per person per day), depends strongly on achieving the other two.
The aim of this paper is multiple. First, to provide a logically-phased overview of the
Silent Revolution over time, paying attention to the key indicators involved; second, to outline
the many socio-economic benefits groundwater irrigation has triggered in arid and semi-arid
countries worldwide; third, to explain how the problems that arise from this Silent Revolution
can be avoided or mitigated and sustainable groundwater development achieved; and fourth,
to show how the lack of awareness on the reality of the Silent Revolution may induce a series
of social and political conflicts (bridging the gap between scholars and decision makers, and
those millions of poor farmers who fulfil their everyday needs with groundwater is an urgent
and necessary action that can be implemented in a fast and economic manner, although it
requires a political willingness).
It must be noted that irrigation nowadays amounts to about 70% of the total world water use
(a figure which increases to about 80–95% in arid and semi-arid countries) and produces about
half the food and fibres required by the human being (United Nations, 2003). Therefore, this
chapter deals in particular with groundwater and agriculture, whilst those problems related to
water in urban areas (or for industry or hydropower) are considered less important in the face
of a potential water crisis, and thus remain beyond the scope of discussion.

2 THE SILENT REVOLUTION OF INTENSIVE GROUNDWATER USE FOR


IRRIGATION

The last decades have witnessed a spectacular increase in the availability of new technologies
such as the submersible pump, as well as significant improvements in the fields of drilling
techniques and hydrogeology. Thus, in arid and semi-arid countries, groundwater irrigation
has become much more profitable de-facto than surface water irrigation. As a consequence, a
real Silent Revolution has taken place: during this period millions of independent (and usually
poor) farmers in these countries have of their own accord implemented the necessary means to
irrigate their land with groundwater. The participation of government agencies in the planning
and control of these groundwater developments has usually been scarce.
Significance of the Silent Revolution of intensive groundwater use 165

Figure 1. Rough (ground)water policy trends in arid and semi-arid countries.

2.1 Stages of the silent revolution


Intensive groundwater development can be divided into a sequence of five stages that take
place over time: hydroschizophrenia, silent revolution, farmer lobbies, conservation lobbies
and social conflict (see Figure 1). It must be noted that while these stages typically span
between 25 and 30 years (the approximate equivalent to one generation), this time might be
longer in some cases, and some overlapping may also occur. For instance, in some countries
where the last stage of social conflict has been reached (like Spain or India), it could be argued
that hydroschizophrenia continues to take place.

2.1.1 Stage 1: Hydroschizophrenia


The first stage corresponds to hydroschizophrenia, a term mentioned by American hydrologist
Raymond Nace in order to describe the attitude of many water decision makers who deal with
surface and groundwater resources separately, often playing down the role of the latter. During
this period, water policy is almost exclusively concerned with large surface water infrastruc-
tures heavily subsidised with public funds. At the same time the potential of groundwater
resources tends to be overlooked.
Since hydrogeology is a fairly recent addition to the natural sciences, governmental water
agencies have traditionally been managed by hydraulic engineers with a significant profes-
sional bias towards large surface water infrastructures. In fact, just about all major hydraulic
works within the last one-hundred years have been under financial and operational control
of government institutions. While this is understandable, unethical attitudes such as arro-
gance, negligence, vested interests, neglect and corruption have often played a role (Llamas &
166 M.R. Llamas & P. Martínez-Santos

Martinez-Santos, 2005). The practical consequence of this is that hydrological plans often
consider surface waters as the only viable resource, even if competent studies prove local
aquifers to be more advantageous, both from the economic and environmental point of view
(Custodio, 2000, 2002). It is important to note that in certain situations, like in some Western
USA states (e.g. California), hydroschizophrenia has been the consequence of legal con-
strains, rather than an attitude as such (the existing regulations were quite different in the case
of surface water and groundwater).
Whilst it is not the aim of this paper to deny the interest of large surface water infrastructures,
it is necessary to realise that these do not always constitute the unique, nor the best solution to
cope with the need of a higher food output. In fact, it will be shown that groundwater irrigation
can generally achieve more significant benefits than traditional surface water systems. It is
also acknowledged that conjunctive use of surface and groundwater resources is in many cases
the ideal solution. Many successful examples of conjunctive use exist in scientific literature,
like the Barcelona water supply (Llamas, 1969), or others in Spain and the USA (Sahuquillo &
Lluria, 2003). However, even if the technical solutions for the implementation of conjunctive
use are well known, the main difficulties often arise from economic, legal and institutional
constrains.
Some places, like the Punjab in Pakistan and India, or the California experience, have
become an example of soil water logging due to over-irrigation from surface water systems.
This problem, caused by the lack of a hydrogeological assessment, may be solved by depleting
the water table through pumping (pumped groundwater can even be used for irrigation if
quality allows). This is an example of poor conjunctive use.

2.1.2 Stage 2: Silent revolution


Wide availability (and affordability) of drilling and pumping techniques has led millions of
farmers to carry out the Silent Revolution in many arid and semi-arid countries worldwide
(Llamas & Martinez-Santos, in press; Deb Roy & Shah, 2003; Moench, 2003; Mukherji,
2003). While exceptions exist (California, Arizona and Texas, for instance), most of these
farmers are poor and often illiterate. Thus, the intrinsic advantages of groundwater develop-
ment, namely resilience against drought and ready availability of the resource on demand,
constitute for them an important incentive for development. In fact, these result in the uncer-
tainties about potential drought impacts being removed from the farmer’s perception. Thus a
new encouragement exists for them to invest in more efficient agricultural methods, including
modern techniques such as sprinkler or drip irrigation.
Security against drought presents further effects. These are, for instance, described by
Chadha (2004a), director of the Groundwater Central Board in India:

“Development of groundwater has led to increased drought proofing of India’s agricultural


economy. The importance of this in the Indian context can be gleaned from the impact of
droughts. In the 1960s groundwater was a relatively insignificant source of irrigation, partic-
ularly in Eastern India. In 1965–1966, rainfall (June to September), was 20% below normal,
leading to drought conditions. National food grain production declined 19% over the previous
year’s level. In contrast, in 1987–1988 rainfall dropped almost 18% below normal, while food
grain production only declined 2% over the previous year’s level. Although the droughts are
not directly comparable, the decline in production in 1987–1988 was significantly smaller
than in 1965–1966, and much of it can be attributed to the spread of irrigation in general and
of groundwater irrigation in particular.
Significance of the Silent Revolution of intensive groundwater use 167

Droughts have ripple effects throughout the Indian economy. Not only is there the direct
loss of production; there are also numerous secondary effects. Vulnerable populations are
particularly at risk, and are often forced to migrate in search for work. Public expenditures
on drought irrigation and food distribution programs also increase substantially. The growth
in India’s irrigated areas, particularly the area irrigated with groundwater, has greatly reduced
the economy’s vulnerability to sharp reductions in rainfall, drought proofing and the rural
economy in general and the crop sector in particular”.

During this stage, the dramatic boost in pumpage may be coupled with a noticeable depletion
of the water table, which in turn results in an increase in abstraction cost per unit volume. In
contrast, the implementation of more efficient irrigation systems offsets this reality, slowing
down the increase in the overall irrigation cost per hectare. Average crop value also begins
to increase, due to the switch to cash crops. Finally, poor farmers become a strong middle
class and are able to send their children to school and to college. As these are often trained
for disciplines other than agriculture, a forward social transition begins to take place towards
industry and the services sector (Moench, 2003).

2.1.3 Stage 3: Farmer lobbies


Ruthless groundwater development takes place in the pursuit of the aforementioned socio-
economic benefits. Consequently, the trends described under the previous heading continue.
Farmers have by then become wealthier and better educated, and as pumping costs increase
with water table depletion, they begin to form strong pressure groups to lobby for perverse
subsidies in the shape of cheap energy and/or water (dams and inter-basin transfers to be
funded by all tax payers).
It is during this stage when groundwater related problems may begin to occur, thus
accelerating the action of farmer lobbies. Such problems include water table depletion,
groundwater quality degradation, land subsidence and ecological impacts upon wetlands and
streamflows, and have been described in Llamas & Martinez-Santos (2005). Water table deple-
tion is the most frequent, and thus there exists a widespread misconception that considers it
the most serious of the lot. However, it may in many cases prove irrelevant (as will be shown
further ahead). In fact, the authors are not aware of any cases where socio-economic havoc has
been triggered by intensive groundwater exploitation of medium to large aquifer systems (sur-
face over 500 km2 ). Moreover, even in very small aquifers, like the case of Crevillente in Spain
(90 km2 ), abstraction takes place from a depth of about 500 m (see Table 2, in section 4.1).
Human-induced water quality degradation, or the hydrocide, is in fact the most significant
one of the problems listed above, since it severely affects crop value and is likely to appear
before water table depletion becomes an issue. However, it must be borne in mind that most
groundwater-quality related problems are usually a consequence of poor land use. For instance,
in the case of many countries in Central and Western Europe, groundwater quality degradation
is not due to intensive pumping, but to the agrochemicals used in rain-fed agriculture.

2.1.4 Stage 4: Conservation lobbies


Large hydraulic infrastructures are often met with controversy, as they tend to demand large
investments of public funds and to clash with the interests of different groups: for instance
environmental organizations (WWF, Greenpeace) and other regional groups who consider
these projects harmful to their rights and interests. These usually unite into a common
front in order to oppose government authority and the economic power of large construction
168 M.R. Llamas & P. Martínez-Santos

companies. Further hostility to these projects may come from scientific and economic-based
analyses, usually carried out by scholars. For instance, the final report of the World Commis-
sion on Dams (WCD, 2000) notes that the economic advantages of many large surface water
irrigation projects are doubtful.

2.1.5 Stage 5: Social conflict


The importance of water resources in the livelihoods of many people, as well as the added
emotional factor which opposing parts may try to exploit in their own favour, can make these
conflicts extend to wide sectors of society. The Ebro River Transfer in Spain, the CALFED
controversy in the USA, or the social unrest due to the Narmada Valley Project provide current
examples of these conflicts – which practically never in historical terms have become real
water-wars (Asmal, 2000).
However, it often seems to be ignored that poor groundwater resources planning is at the
very root of these conflicts. S.D. Limaye (July 2004, pers. comm.), vice-president of the IAH
(International Association of Hydrogeologists) Asian Chapter, explained this situation in the
following terms:
“The attitude of major international funding agencies and other international organizations,
towards groundwater as a resource and also towards countries making extensive use of ground-
water, has always been skeptical. For example: World Bank and Asian Development Bank
have sponsored several projects for surface water development. Their support for groundwater
projects has been minimal, on the grounds that it is a hidden resource, difficult to explore and
assess, energy is required for pumping, . . .
In developing countries, politicians have always preferred surface water projects because
they are grand, big-budget items with a great impact on the people in their constituencies and
a greater opportunity for corruption”.
As long as this situation continues, the current gap between the way millions of farmers
worldwide attend to their everyday water needs and the perception of water policy decision
makers seems likely to remain.

2.2 Key indicators of intensive groundwater use


In Figure 1, a series of social, economic and technical indicators (namely, crop value, abstrac-
tion cost, depth to water table, irrigation cost and % population in agriculture), constitute the
key indicators to follow in the Silent Revolution of intensive groundwater development. A
somere description of these is provided below.
(a) Groundwater abstraction cost per unit volume (US$/m3 ): Groundwater abstraction cost
increases with depth (because more energy is required for pumping and the need to drill
deeper wells). Abstraction costs usually range between US$ 0.01/m3 and US$ 0.20/m3 .
Presently, the most significant water table depletion known to the authors has taken place
in Crevillente (Spain), where groundwater is being pumped from a depth of 500 m at a cost
of US$ 0.30/m3 . See Hamer (2002), for California; Chebaane et al. (2004), for Jordan;
Hernández-Mora et al. (2001), for Spain; Deb Roy & Shah (2003), for Southeast Asia.
An extreme case is the cost of groundwater abstraction carried out by one-dollar-a-day
farmers by means of treadle pumps (Polak, 2004).
(b) Depth to water table (m): Intensive groundwater use causes a depletion rate in the water
table often ranging between 0.5 and 5 m/yr (even more in some cases). However, this is
Significance of the Silent Revolution of intensive groundwater use 169

seldom a deterrent issue except in shallow aquifers, which could reach physical exhaustion.
In fact, even the increase in the cost of energy for pumping is a relatively small problem in
comparison with potential groundwater quality degradation (like for instance in the case
of seawater intrusion in coastal aquifers), and certain equity issues arising from the drying
of shallow wells (usually owned by poor farmers who may not have the possibility to drill
any further).
(c) Irrigation cost (US$/ha): Groundwater abstraction cost per unit volume may increase with
time (up to ten-fold due to water table depletion). While irrigation cost also increases, it
does so at a lower rate, as farmers begin to use more efficient agricultural and irrigation
technology and switch to lower water-consuming crops (for instance, from corn or rice to
grapes and olive trees). It is estimated that irrigation cost ranges between US$ 20/ha (e.g.
in the case of cereals in good shallow aquifers) and US$ 1000/ha (in aquifers where water
is pumped from a depth of hundreds of metres and crop water consumption is up to 3000
or 4000 m3 /ha/yr).
(d) Population in agriculture (%): This is a measure of the social transition triggered by
groundwater development. Groundwater irrigation usually helps to improve the socio-
economic status of poor farmers to a greater extent than traditional surface water systems
(Deb Roy & Shah, 2003). This is a consequence of the entrepreneurial mentality of those
farmers who decide to invest in their own wells, as well as to crop security due to the
resilience of aquifers against drought. In the meantime, those farmers have been able
to send their children to college to receive a better education. Some of these will be
trained in agricultural disciplines, and then return to farming (albeit with a better grasp
for technology). Others, however, will move on to become engineers, teachers, doctors,
economists and so forth. As a consequence, within one or two generations, farmers have
a chance to move from agriculture onto industry and the services sector. In Spain, for
instance, the share of population in agriculture has decreased from 50% to below 6% in
the last two generations. Similar examples can be found in many industrialised countries,
where the main economic sector a century ago was probably agriculture (the case of the
USA, a country where 3% of the population is able to feed the remaining 97%, and which
still is the most important food-exporter in the world).
(e) Crop value (US$/ha): This is perhaps the most significant aspect of the Silent Revolution.
As farmers become richer and more educated, they move from low-value to cash crops. This
is mainly due to the ready availability of groundwater and the resilience of aquifers against
drought, which also encourage farmers to make greater investments in order to improve
their agricultural and irrigation technology and to grow cash crops. Crop value is not
only related to crop type. In fact, the value of cash crops per hectare also depends largely
on agricultural techniques applied and market conditions. Thus, it ranges widely: for
instance, between US$ 500/ha (e.g. cereals) and over US$ 70,000/ha (tomatoes, cucumber
and other greenhouse crops sold in other European states) in Spain, although these figures
are probably similar in many other countries.

It must be noted that these indicators do not take into account the cost of externalities such
as the degradation of aquatic ecosystems. There are two main reasons for it: one, the difficulty
of assessing such cost, as described by the National Research Council (1997); and two, the
fact that, according to Kuznets environmental curve, poverty is really the worst enemy of the
environment (a good example of this is the situation in Southeast Asia). It must be borne in
170 M.R. Llamas & P. Martínez-Santos

mind that concepts such as sustainability or the value of ecosystems are not easy to understand
for governments nor for one-dollar-a-day farmers in developing countries. The situation is very
different in richer countries, where the socio-economic setting is radically different, and a series
of different mechanisms exist to protect the environment (for instance the Western USA, where
an important tool for the defence of water environments is the Endangered Species Act).

3 GROUNDWATER AND HUMAN DEVELOPMENT: SOME BENEFITS OF THE


SILENT REVOLUTION

3.1 Socio-economic benefits of intensive groundwater use


As stated before, the Silent Revolution is a market-driven phenomenon. In fact, the socio-
economic efficiency of groundwater irrigation is between 3 and 10 times higher than that of
surface water systems in terms of economic productivity (US$/m3 ) and employment generation
(jobs/m3 ). Table 1 shows a socio-economic assessment of these figures in Andalusia, Spain.
This analysis has been extended to other regions of Spain (Hernández-Mora & Llamas, 2001)
with different climatic and social conditions, and the results are also similar. A recent paper
(Vives, 2003) seems to confirm that the trend continues.
Data from countries such as India provide even more spectacular results. According to
Dhawan (1995), yields in areas irrigated with groundwater are 1/3 to 1/2 higher than those in
areas irrigated with surface resources. In a previous report, Dains & Pawar (1987) estimated
that as much as 70–80% of India’s agricultural output might be groundwater dependent. The
Indian Water Resources Society (1999) published, among others, the following significant
data:
– Groundwater is contributing at present 50% of irrigation surface, 80% of water for domestic
use in rural areas, and 50% of water in urban and industrial areas.
– Groundwater abstraction structures had increased from 4 million in 1951 to nearly 17 million
in 1997.
– In the same period the groundwater irrigated area had increased from 6 to 26 million ha,
and it is estimated that this rapid pace of development is likely to continue and will reach
64 million ha in the year 2007.
By indirect calculation it may be estimated that in India the average amount of water applied
in surface water irrigation is around 16,000 m3 /ha/yr; in groundwater irrigation this ratio is

Table 1. Comparative socio-economic efficiency of surface water and groundwater


irrigation in Andalusia, Spain.

Indicator Surface water Groundwater Total

Irrigated surface (103 ha) 600 210 810


Average economic production (a/ha) 3300 8600 4600
Average consumption at origin (m3 /ha/yr) 7400 4000 6500
Water productivity (a/m3 ) 0.42 2.16 0.72
Employment generated (EAJ*/106 m3 ) 17 58 25

*EAJ: Equivalent annual jobs


Source: Hernández-Mora et al. (2001).
Significance of the Silent Revolution of intensive groundwater use 171

only 4000 m3 /ha/yr. In other words, it seems that in India and on average, the economic yield
in irrigation and by m3 is from 5 or 10 times higher when groundwater is used than when
irrigation is carried out with surface water.
Some recent reports indicate that the spectacular increase of groundwater use in Southeast
Asia continues. Chadha (2004a, 2004b), director of the Groundwater Central Board, considers
that the renewable groundwater recharge in India is in the order of 450 km3 /yr. In 1997, surface
irrigated was about 46 million ha, and the total number of wells about 17 million. According to
the same source, total pumping amounted to 150 km3 /yr, and the total usable groundwater stor-
age (down to 450 m in alluvial aquifers and 150 m in hard rock aquifers) to 11,000 km3 . A few
months after the previous reports, Shah (in press) points out that 60% of India’s irrigated area
is served by groundwater wells. An independent survey suggests that the figure may well be
75%, and even more if conjunctive uses are included. Shah estimates that India has been adding
0.8–1 million new tubewells every year since 1996, and that current pumping volume is in the
order of 200 km3 /yr (approximately twice the figure quoted by the Indian Water Resources
Society).
Overall, most authors (almost without exception) conclude that groundwater irrigation is a
major catalyst for rural development and that its productivity, even in the case of farmers who
do not have a well and need to purchase water off others, is higher than that of surface water
irrigation.
What are the reasons for the higher productivity of groundwater irrigation in relation to
traditional surface water systems? The answer to this question, as previously stated, is multiple:

(a) In the first place are the physical advantages of groundwater in terms of drought resilience
and ready availability of the resource on demand. These result in crop security, which
in turn leads farmers to invest in better agricultural and irrigation technologies. Since
abstraction cost (US$/m3 ) is higher than that of traditional surface water systems, wasteful
over-irrigation is discouraged (however, when electrical energy is heavily subsidised, the
incentive for efficient irrigation is smaller).
(b) Due to the above, farmers increasingly tend to a switch towards cash crops. A more efficient
use ensues, thus maximising economic return per m3 . In fact, while pumping cost increases
with time (due to water table depletion), so does crop value, but at a much higher rate.
Overall, it appears that crop value tends to be at least 20 times higher than abstraction cost.

The available data shows that the above seems to hold in arid and semi-arid regions world-
wide. Therefore, there is a real need to carry out world-scale analysis of these figures in
order to: (1) challenge those so-called world water crises; (2) give groundwater its rightful
importance within the framework of global water policy; and (3) avoid problems to come if
the current situation of chaotic groundwater development is to continue.

3.2 Intensive groundwater use as a catalyst for social transition


Since groundwater is a resource available and accessible to the poor, it has often acted as a
catalyst for a quick social transition: often from a poor and illiterate society to a well-developed
middle class, with a strong industry and/or services sector (Moench, 2003).
Several documented cases of this phenomenon exist in the scientific literature. For instance,
Chebaane et al. (2004) cite groundwater-based irrigation development as the main driving
force behind small villages becoming important provincial capitals in Jordan.
172 M.R. Llamas & P. Martínez-Santos

It must be acknowledged that a series of tragic examples exist in reference to suicides by


Indian farmers (due to well failure in poor hard-rock aquifers), as well as to the re-location
of small communities in Yemen – as a consequence of water table depletion, according to
Moench (2003). While these are indeed sad cases (and cannot be overlooked), it is important
to note that they correspond to very isolated situations under very specific hydrogeological
settings. In fact, the authors of this chapter do not know of any cases of medium or large
aquifers (surface area larger than 500 km2 ) where socio-economic havoc has been triggered
by intensive groundwater use.

3.3 Groundwater developments are less prone to corruption than large surface water
infrastructures
Groundwater development is, by its own very nature, generally less prone to corruption than
large surface water infrastructures. Indeed, the very nature of the latter (often large in size
and investments) makes them a juicy prospect in the eyes of corrupt officials and corporate
decision-makers. On the other hand, groundwater developments are usually carried out by
individuals whose economic means are scarce.
Cases of corruption are not of an isolated nature. In a few international conferences and
reports the practical importance of corruption and bribery in the water management has been
emphasised. This was an issue that previously was very rarely mentioned in such conferences
or in international reports. Perhaps the most important document on this topic is the book by
OECD (2000) dealing with the Convention on Combating Bribery of Foreign Public Officials
in International Business Transactions. This book was specifically quoted in the Report of the
World Commission on Dams (WCD, 2000: 186–187, 249). This author has also mentioned the
practical relevance of corruption as a frequent driving force in water policy in several papers
(Delli Priscoli & Llamas, 2001; Llamas & Delli Priscoli, 2000; Llamas & Custodio, 2003;
Llamas & Martínez-Santos, 2005). Postel (1999: 229) presents an interesting view of the role
of consulting firms, construction companies and politicians in promoting large surface water
irrigation projects, which is transcribed below:
“The rules, especially for large schemes, often look something like this: a politician seizes
the potential for bolstering political support by proposing an irrigation project in a strategic
location. Engineering firms lobby the decision-makers in order to raise their chances of win-
ning the project’s construction contract. The politician, in collusion with colleagues who also
want to please their constituents with pork-barrel projects, sees to it that the nation’s taxpayers
pay most of the bill. The farmers themselves pay only a small fraction of the project’s cost; in
return, they support the politician in the next election. With prices kept artificially low, farmers
have little incentive to use water efficiently. The irrigation system never becomes financially
self-sustaining because the meager fees collected from the farmers do not cover the system’s
operation and maintenance costs, much less its capital costs. National or state irrigation agen-
cies keep the projects running with taxpayer funds. If budgets become tight and maintenance
work is neglected, the systems fall into disrepair. Gradually, agricultural output and benefits
begin to decline. Either the government saves the project by allocating more taxpayer money
for expensive rehabilitation work, or the system deteriorates until farmers abandon it. Alterna-
tively, international donors come to the rescue with funds provided by taxpayers in wealthier
countries”.
Another issue to be considered is the almost universal policy of public perverse subsidies
for water and agriculture. According to Myers & Kent (1998), these subsidies are those which
Significance of the Silent Revolution of intensive groundwater use 173

are noxious both for the economy and the environment. In most cases, the water users only
pay a small fraction of the real cost of the water supplied. This is especially true in surface
water for irrigation. Water policy all over the world has, during the past decades, focused on
the management of the supply and not on the management of the demand. This has induced
an almost universal wasteful use of water.
In most groundwater developments the situation may be quite different. Owners of wells
usually pay for their construction, maintenance and operation. But they do not usually pay the
external costs caused by the impact of groundwater abstraction.

4 NEGATIVE IMPACTS ARISING FROM INTENSIVE GROUNDWATER


DEVELOPMENT

It would be unwise to advocate widespread groundwater use as a universal solution for the
world’s water problems. Uncontrolled groundwater development can lead to a series of effects
including water table depletion (the most frequent), groundwater quality degradation, land
subsidence as well as to other environmental impacts such as streamflow reductions and
desiccation of wetlands. As a consequence, certain water officers (and often journalists) have
voiced the hydromyth of groundwater as a fragile resource.
Out of these impacts, groundwater quality degradation, wisely termed hydrocide by Prof.
Lundqvist, is the most significant one, albeit not the most frequent. In fact, the huge storage
capacity of most aquifers suggests that it usually takes about 2–3 generations of heavy pumping
before any of such negative effects becomes significant. In any case, it can be shown that most
of these problems are due to poor groundwater governance and land use planning, rather than
to intensive groundwater use.

4.1 The controversy on groundwater mining


Groundwater mining (pumping at a rate in excess of aquifer recharge) is often seen by many as a
kind of ecological sin, opposed to the basic tenets of sustainable development. Nevertheless, a
good number of authors disagree with this view, considering that under certain circumstances,
groundwater mining can be a sensible alternative (Freeze & Cherry, 1979; Issar & Nativ, 1988;
Llamas, 1992, 1999; Collin & Margat, 1993; Margat, 1994; Lloyd, 1997; Custodio, 2002;
Price, 2002; Abderrahman, 2003).
Sustainability is often defined as “meeting the current needs without compromising those
of future generations”. In this regard, it is necessary to clarify that this definition is not
comprehensive, and that sustainable development does not constitute an end in itself (Price,
2002). Indeed, questions such as “will future generations need it more than the present one?”,
do not find an answer in this definition.
Whether groundwater mining falls under the umbrella of sustainability has to be dealt with
on a case-by-case basis. In most countries it is considered that groundwater abstraction should
not exceed the renewable resources. In other countries – mainly in the most arid ones – it
might be considered that groundwater mining is an acceptable policy, as long as available data
assure that the groundwater development can be economically maintained for a long time,
for example, more than fifty years and that the potential ecological costs and socio-economic
benefits have been adequately evaluated.
174 M.R. Llamas & P. Martínez-Santos

In Saudi Arabia, according to Dabbagh & Abderrahman (1997), the main aquifers (within
the first 300 m of depth) contain huge amount of fresh fossil water (a minimum of 2000 km3 )
that is 10,000 to 30,000 years old. It is considered that these fossil aquifers can supply useful
water for a minimum period of 150 years. Current abstraction seems to be around 15–20 km3 /yr.
During a couple of decades the Saudi government has pumped several km3 /yr of non-renewable
groundwater to grow low cost crops (mainly cereals) which were also heavily subsidised. The
official aim of such activity was to help to transform nomadic groups into farmers. Apparently
such overdraft has been a success. Now the amount of groundwater abstraction has been
dramatically reduced and the farmer nomads have become high-tech farmers growing cash
crops. Another example is the situation of the Nubian sandstone aquifer located below the
Western desert of Egypt. According to Idriss & Nour (1990), the fresh groundwater reserves
are higher than 200 km3 and the maximum pumping projected is lower than 1 km3 /yr. Probably
similar situations do exist in Libya and Algeria. Other examples of mining groundwater can
be found in Llamas & Custodio (2003).
The Indian case has a special interest in this regard, not only because India is the number
one country in the world when it comes to groundwater use, but also because a good number
of authors have dismissed groundwater development as a pillar of sand, soon to collapse. A
recent paper by the director of the Indian Central Groundwater Board (Chadha, 2004a) states
that the number of dark blocks (by blocks meaning aquifers) in India has grown from some 250
in 1984–1985 to 450 in 1997–1998. However, this author puts a question mark on the World
Bank’s categorisation of blocks into white, grey and dark. There is not enough space here to
discuss the manifold concept of overexploitation (see Custodio, 2002; Llamas & Custodio,
2003). Yet, it should suffice to say that such categorisation in blocks is based on the estimated
difference between pumping and recharge, when it is well known that groundwater recharge
can be substantially increased by water table depletion. It seems, in any case, that the accuracy
of the data used in this assessment is rather low.
It is not easy to achieve a virtuous middle. As Collin & Margat (1993) state: “we move
rapidly from one extreme to the other, and the tempting solutions put forward by zealots
calling for Malthusian underexploitation of groundwater could prove just as damaging to the
development of society as certain types of excessive pumping”.
An extreme example of groundwater mining is the Crevillente aquifer, Spain (Table 2), a
small (90 km2 ) limestone formation where heavy pumping has taken place since the 1970s. As
a consequence, the water table has been depleted at a rate of 10–15 m/yr, and required pumping
elevation now reaches 500 m (wells are up to 700 m deep). The approximate abstraction cost

Table 2. Groundwater development in the Crevillente aquifer, Spain:


an extreme case.
Aquifer settings (size, geology) 90 km2 , limestones
Estimated recharge/abstraction 2–16 Mm3 /yr
Initial pumping elevation (1970s) 20–30 m
Current pumping elevation 500 m
Groundwater abstraction cost 0.30 a/m3
Irrigation cost (grapes) 1000 a/ha/yr → 3300 a/ha/yr
Crop value 25,000 a/ha → 15,000 a/ha

Source: F. Corchón, of the Jucar Basin Authority (June 2004, pers. comm.).
Significance of the Silent Revolution of intensive groundwater use 175

is currently 0.30 a/m3 . While these world-record figures are in place, the system has not yet
collapsed.
However, crop value has gone down from 25,000 a/ha to about 15,000 a/ha, due to the
lower fertility of the soil associated with an increase in groundwater salinity (not so much
with the extra cost induced by groundwater depletion). This has become a pressing issue, and
thus the Crevillente farmers’ lobby, together with others (farmers who obtain their water from
small nearby aquifers as well as urban water supply companies) have obtained the approval of
a seemingly needless inter-basin transfer (a total of 90 Mm3 /yr with a cost of 230 million a,
to be funded 2/3 by public funds and 1/3 by the beneficiaries).
The idea that groundwater mining might be an ethically sound practice under certain cir-
cumstances was developed by Llamas (1999) in a UNESCO Conference held in Lybia in 1998.
This thesis has been followed in the first report of the Chairman of the Subcommittee of the
UNESCO World Commission on the Ethics of Science and Technology (Selborne, 2000; Delli
Priscoli & Llamas, 2001).

4.2 Water quality degradation: the hydrocide


As stated above, water quality degradation is the main problem associated with intensive
groundwater use. Groundwater abstraction can certainly cause, directly or indirectly, changes
in groundwater quality. The change in hydraulic gradient due to groundwater abstraction might
constitute cause of quality degradation, as it may result in intrusion from adjacent saline water
bodies (other aquifers or the sea). This is a typical problem in many coastal regions in arid
and semi-arid regions, although it is well known that coastal aquifers, like the ones in Israel
or in Orange County, California, can be managed in a sustainable way. The relevance of
the saline water intrusion not only depends on the total withdrawal in relation to the natural
groundwater recharge, but also on the well field location and design, and on the geometry
and hydrogeological parameters of the aquifer. Thus, in many cases the existing problems are
due to uncontrolled and unplanned groundwater development and not to excessive pumping
(Custodio & Bruggeman, 1982).
Quality degradation may not be related at all to excessive groundwater abstraction in relation
to average natural recharge. Other causes may be responsible, such as return flow from surface
water irrigation, leakage from urban sewers, infiltration ponds for wastewaters, septic tanks,
urban solid waste landfills, abandoned wells, mine tailings and many other activities (Barraqué,
1997).

4.3 Land subsidence


Sedimentary formations are deposited at low density and large porosity. As subsequent layers
are deposited the overburden compresses the underlying strata. The overburden is in static
equilibrium with the intergranular stress and the pore water pressure. This equilibrium is
quickly reached in coarse-granular layers, but in fine-grained layers with low permeability it
may take a long time. The effect of this process is the natural progressive consolidation of
sediments.
When an aquifer is pumped the water pore pressure is decreased and the aquifer solid
matrix undergoes a greater mechanical stress. This greater stress may produce compaction of
the existing fine-grained sediments (aquitards) if the stress due to the decrease in water pore
pressure is greater than the so-called preconsolidation stress. This is a well-known situation
176 M.R. Llamas & P. Martínez-Santos

which has occurred in some aquifers formed by young sediments, such as those in Mexico
City, Venice, Bangkok and others (Poland, 1985). The Mexico City case is probably the most
dramatic one, although since it is related to urban water supply and not to irrigation, it will
not be discussed here.
Caves and other types of empty spaces may exist under the water table in karstic aquifers.
When the water table is naturally depleted the mechanical stability of the roof of such empty
spaces may be lost and the roof of the cave collapses. This is a natural process that gives rise to
the classical dolines and poljes in the karstic landscape. When the water table depletion or oscil-
lation is increased by groundwater abstraction, the frequency of karstic collapses can be also
increased. The accurate prediction of such collapses is not easy (LaMoreaux & Newton, 1992).
In both cases the amount of subsidence or the probability of collapses is related to the
decrease in pore water pressure which is related to the amount of groundwater withdrawal.
Nevertheless the influence of other geotechnical factors may be more relevant that the amount
of water abstracted in relation to the renewable groundwater resources of the aquifer.

4.4 Environmental impacts on wetlands and streamflows


The ecological cost of groundwater development should be compared with the socio-economic
benefits produced (Barbier et al., 1997; National Research Council, 1997). The evaluation of
the ecological impacts is highly dependent on the social perception of ecological values in the
corresponding region, a perception which is changing rapidly in most countries. For instance,
the Framework Directive on Water of the European Union of the year 2000 paid attention to
monitoring and conservation of aquatic ecosystems and especially to wetlands.
In arid and semi-arid regions, wetlands or oases are usually rare and frequently related to
groundwater discharge zones. The development of groundwater for irrigation or other uses may
often have a significant negative impact on the hydrological functioning of wetlands or oases
(Fornés & Llamas, 1999). These impacts should be properly evaluated by decision-makers,
bearing in mind that there is no blueprint solution. This issue, very much in relation with
Kuznets environmental curve, is particularly relevant in the case of developing countries, where
poverty constitutes the worst enemy of the environment, and where paradigms proper to wet
industrialised countries are not likely to succeed. On the other hand, in industrialised countries
such as Spain, conflicts between farmers and conservation groups are relevant (Brufao &
Llamas, 2003; Llamas, 2003).

5 STRATEGIES FOR SUSTAINABLE GROUNDWATER USE IN ARID AND


SEMI-ARID REGIONS

There exists a general consensus that, in order to avoid conflicts and to move from confrontation
to cooperation, water development projects require the participation of the social groups
affected by the projects, the stakeholders. The participation should begin in the early stages
of the project and should be, as much as possible, bottom-up and not top-down. The first
question is to define who the stakeholders are; the second, how, when and where they should
intervene in the decision making processes.
The Spanish experience, in trying to implement groundwater management as a public
dominion, indicates clearly that the active collaboration of Groundwater Users Associations is
Significance of the Silent Revolution of intensive groundwater use 177

a key element (Hernández-Mora & Llamas, 2000, 2001). However, the process implementation
demands sometime to switch from old to new paradigms (Lopez Gunn & Llamas, 2001).
In July 2001, the Spanish Parliament approved the Law of the National Water Plan. This
Plan included several provisions that, if really enforced, would change the chaotic situation of
groundwater development in Spain. Perhaps the most important article of this Law was the one
which strictly demanded setting up of Groundwater Users Associations in every intensively
developed aquifer and a thorough hydrogeological assessment of every aquifer which may be
supposed to receive a surface water transfer from other catchments. The National Water Plan
Law also stated that an intense and broad Water Education Programme has to be implemented.
These provisions, together with the 1000 Mm3 /yr Ebro River Transfer to the overexploited
aquifers of Southeastern Spain, have however been changed due to the government overturn of
March 2004. The new Plan, subject to a heated economic and environmental debate, advocates
the building of a series of desalination plants in the Mediterranean coast. As it can be easily
seen, the paradigms from a century of perverse subsidies have not been overcome, and the
existing administrative chaos in relation to groundwater seems not to be in the mind of Spanish
water policy decision-makers.
Obviously, there is no blueprint for a universal solution. For example, in some arid and
semi-arid developing countries, when dealing with correction of ecological impacts of over-
exploitation, the influence of conservationists groups will probably be weak compared to the
influence of farmers associations or urban water supply companies.
The necessary participation of the stakeholders demands that they are aware of the way
the issue at hand will affect them directly or indirectly, and also a basic knowledge of the
hydrogeological concepts involved in aquifer development. Probably in most countries there
exist a good number of hydromyths or obsolete paradigms about the origin, movement and
potential pollution of groundwater. In any stressed aquifer it is essential to organise different
types of educational activities aimed at different groups: from school students and teachers
to officials of Water Administrations, as described by McClurg & Sudman (2003), and other
authors and institutions.

6 CONCLUSIONS

In the last few decades, a Silent Revolution of intensive groundwater use for irrigation has
taken place in arid and semi-arid countries worldwide. This has been carried out by millions
of independent farmers who, at their own expense and risk, have drilled wells in order to
access the intrinsic advantages of groundwater. An analysis of the key social, economic and
technical indicators involved, reveals that the guaranteed crop value is usually much higher
that the pumping cost (at least 20 times, in most cases). Therefore, this Silent Revolution is
market-driven, and thus unstoppable if business continues as usual.
The Silent Revolution has taken place with little or no control on the part of government
authorities. As a consequence a series of problems, at times pretended and at times genuine,
have arisen in some places. These have been often magnified, giving rise to the hydromyth of
groundwater development as a pillar of sand (a fragile resource). However, to the knowledge
of these authors, there are no documented cases where the Silent Revolution has induced
socio-economic havoc. This may be due to the huge storage capacity of most aquifers, which
allows for one or two generations of heavy pumping before serious impacts begin to be felt (it
178 M.R. Llamas & P. Martínez-Santos

must be borne in mind that in the meantime a social transition away from agriculture has also
taken place). However, it seems that this situation cannot last forever.
Therefore, three fairly cheap and concrete steps are suggested towards a sustainable use of
groundwater resources:

(1) Carry out a thorough and transparent assessment of the real situation of surface water
and groundwater, together with their respective technical, economic and social efficiency.
This assessment could probably carried out within a year in most cases.
(2) Implement a massive water education campaign, paying attention to the special character-
istics of groundwater, stressing the need for a common pool resource type development.
(3) Governments should foster the creation of groundwater users associations in order to
manage intensively developed aquifers.

REFERENCES

Abderrahman, W.A. (2003). Should intensive use of non-renewable groundwater resources always be
rejected? In: M.R. Llamas & E. Custodio (eds.), Intensive Use of Groundwater. Challenges and
opportunities. Balkema Publishers. Lisse, the Netherlands: 191–206.
Asmal, K. (2000). Water: from casus belli to catalyst for peace. Stockholm Water Week (address in the
opening session, 14th August 2000).
Barbier, E.B.; Acreman, M. & Knowler, D. (1997). Economic evaluation of wetlands: a guide for policy
makers and planners. Ramsar Convention Bureau. Gland, Switzerland: 127 pp.
Barraqué, B. (1997). Groundwater management in Europe; regulatory, organizational and institutional
change. Proceedings of the International Workshop: how to cope with degrading groundwater quality
in Europe. Stockholm, Sweden, 21–22 October: 16 pp (preprint).
Brufao, P. & Llamas, M.R. (eds.) (2003). Conflictos entre el desarrollo de las aguas subterráneas y
la conservación de humedales: aspectos legales, institucionales y económicos. Fundación Marcelino
Botín and Ediciones Mundi-Prensa. Madrid, Spain: 337 pp.
Chadha, D.K. (2004a). Status of groundwater development and management in India. In: Resources
Conservation and Food Security. Concept Publishing Co., New Delhi, India. Vol. 1: 9–30.
Chadha, D.K. (2004b). Groundwater potential development and population: a critical review. In:
Sundaram et al. (eds.), National Resources Management and Livelihood Security. Concept Publishing
Co., New Delhi, India: 42–77.
Chebaane, M.; El-Naser, H.; Filch, J.; Hijazi, A. & Jabbarin, A. (2004). Participatory groundwater
management in Jordan: development and analysis of options. Hydrogeology Journal, 12(1): 14–33.
Collin, J.J. & Margat, J. (1993). Overexploitation of water resources: overreaction or an economic reality?
Hydroplus, 36: 26–37.
Custodio, E. (2000). The complex concept of groundwater overexploitation. Papeles Proyecto Aguas
Subterráneas, A1. Fundación Marcelino Botín, Santander, Spain: 58 pp.
Custodio, E. (2002). Aquifer overexploitation: what does it mean? Hydrogeology Journal, 10:
254–277.
Custodio, E. & Bruggeman, G.E. (1982). Groundwater problems in coastal areas. Studies and Reports
in Hydrology, 45. UNESCO, Paris, France: 650 pp.
Dabbagh, A.E. & Abderrahman, W.A. (1997). Management of groundwater resources under various
irrigation water use scenarios in Saudi Arabia. The Arabian Journal for Science and Engineering,
22(IC): 47–64.
Dains, S.R. & Pawar, J.R. (1987). Economic return to irrigation in India. Report prepared by SDR
Research Group Inc. for the U.S. Agency for International Development. New Delhi, India.
Deb Roy, A. & Shah, T. (2003). Socio-ecology of groundwater irrigation in India. In: M.R. Llamas & E.
Custodio (eds.), Intensive Use of Groundwater. Challenges and opportunities. Balkema Publishers.
Lisse, the Netherlands: 307–336.
Significance of the Silent Revolution of intensive groundwater use 179

Delli Priscoli, J. & Llamas, M.R. (2001). International perspective in ethical dilemmas in the water
industry. In: C.K. Davis & R.E. McGinn (eds.), Navigating in Rough Waters. American Water Works
Association. Denver, Colorado, USA: 41–64.
Dhawan, B.D. (1995). Groundwater depletion, land degradation and irrigated agriculture in India.
Commonwealth Publisher. New Delhi, India.
Fornés, J. & Llamas, M.R. (1999). Conflicts between groundwater abstraction for irrigation and wetlands
conservation: achieving sustainable development in La Mancha Húmeda Biosphere Reserve (Spain).
In: C. Griebler et al. (eds.), Groundwater Ecology. A Tool for Management of Water Resources.
European Commission. Environmental and Climate Programme: 227–236.
Freeze, R.A. & Cherry, J.A. (1979). Groundwater. Prentice-Hall. Englewood Cliffs. New Jersey, USA:
604 pp.
Hamer, G. (2002). The cost of water in Southern California. In: Proceedings of California Ground-
water Resources Association Conference (September). Newport Beach. Power Point Presentation,
32 slides.
Hernández-Mora, N. & Llamas, M.R. (2000). The role of user groups in Spain: participation and conflict
in groundwater management. CD-ROM of the X World Water Congress. Melbourne, Australia, 12–16
March. International Association of Water Resources: 9 pp.
Hernández-Mora, N. & Llamas, M.R. (eds.) (2001). La Economía del Agua Subterránea y su Gestión
Colectiva. Fundación Marcelino Botín and Ediciones Mundi-Prensa. Madrid, Spain: 549 pp.
Hernández-Mora, N.; Llamas, M.R. & Martínez Cortina, L. (2001). Misconceptions in Aquifer Over-
exploitation. Implications for Water Policy in Southern Europe. In: C. Dosi (ed.), Agricultural Use of
Groundwater. Towards Integration between Agricultural Policy and Water Resources Management.
Kluwer Academic Publishers: 107–125.
Idriss, H. & Nour, S. (1990). Present groundwater status in Egypt and environmental impacts.
Environmental Geology and Water Sciences, 16(3): 171–177.
Indian Water Resources Society (1999). Water: Vision 2050. New Delhi, India: 74 pp.
Issar, A.S. & Nativ, R. (1988). Water beneath the desert: keys to the past, a resource for the present.
Episodes, 11(4): 256–262.
LaMoreaux, P.E. & Newton, J.G. (1992). Environmental effects of overexploitation in a karst ter-
rain. In: I. Simmers, F. Villarroya & L.F. Rebollo (eds.), Selected Papers on overexploitation.
International Association of Hydrogeologists. Selected Papers, 3. Heise, Hannover, Germany:
107–113.
Llamas, M.R. (1969). Conjunctive use of surface and groundwater in the water supply of Barcelona,
Spain. Bulletin of the International Association of Scientific Hydrology, 14(3): 119–136.
Llamas, M.R. (1992). La surexploitation des aquifères: aspects techniques et institutionnels. Hydroge-
ologie, 4: 139–144. Orleans, France.
Llamas, M.R. (1999). Consideration on Ethical Issues in Relation to Groundwater Development and/or
Mining. In: UNESCO International Conference on RegionalAquifer Systems inArid Zones. Managing
Non-Renewable Resources. Tripoli, Libya: 20–24. Also in: Technical Documents in Hydrology, V
IHP, No. 42, UNESCO, Paris, France: 467–480.
Llamas, M.R. (2003). Epilogue. In: Special Issue on Water in the Iberian Peninsula. Water International,
28(3): 405–409.
Llamas, M.R. & Custodio, E. (eds.) (2003). Intensive Use of Groundwater. Challenges and
Opportunities. Balkema Publishers. Lisse, the Netherlands: 478 pp.
Llamas, M.R. & Delli Priscoli, J. (2000). Water and Ethics. Papeles Proyecto Aguas Subterráneas, A5.
Fundación Marcelino Botín. Santander, Spain: 56–99.
Llamas, M.R. & Martínez-Santos, P. (2005). Ethical issues in relation to intensive groundwater use.
In: A. Sahuquillo, J. Capilla, L. Martínez Cortina & X. Sánchez Vila (eds.), Selected Papers of
the Symposium on Intensive Groundwater Use (SINEX). IAH Selected Papers. Taylor and Francis
Publishers, the Netherlands: 17–36.
Llamas, M.R. & Martínez-Santos, P. (in press). Intensive groundwater use: a silent revolution that cannot
be ignored. In: Proceedings of the 2004 Stockholm World Water Week. Stockholm International Water
Institute, Sweden.
Lloyd, J.W. (1997). The future use of aquifers in water resources management in arid areas. The Arabian
Journal for Science and Engineering, 22(IC): 33–45.
180 M.R. Llamas & P. Martínez-Santos

López Gunn, E. & Llamas, M.R. (2001). New and old paradigms in Spain’s Water Policy. In: Water
security in the Third Millenium: Mediterranean countries towards a regional vision. UNESCO Science
for Peace Series, Vol. 9: 271–293.
Margat, J. (1994). Groundwater operations and management. Groundwater Ecology. Academic Press:
505–522.
McClurg, S. & Sudman, R.S. (2003). Public and stakeholder education to improve groundwater. In:
M.R. Llamas & E. Custodio (eds.), Intensive Use of Groundwater. Challenges and Opportunities.
Balkema Publishers, Lisse, the Netherlands: 271–286.
Moench, M. (2003). Groundwater and Poverty: exploring the connections. In: M.R. Llamas & E.
Custodio (eds.), Intensive Use of Groundwater. Challenges and Opportunities. Balkema Publishers,
Lisse, the Netherlands: 441–456.
Mukherji, A. (2003). Groundwater development and agrarian change in Eastern India. Comment No.
9 in IWMI-TATA Water Policy Program (http://www.iwmi.org/iwmi-tata): 11 pp.
Myers, N. & Kent, J. (1998). Perverse subsidies: their nature, scale and impacts. International Institute
for Sustainable Development. Winnipeg, Canada: 210 pp.
National Research Council (1997). Valuing Ground Water. National Academy Press, Washington D.C.,
USA: 189 pp.
OECD (Organisation for Economic Cooperation and Development) (2000). No longer business as usual:
fighting bribery and corruption. Paris, France: 276 pp.
Polak, P. (2004). Water and the other three revolutions needed to end rural poverty. Invited paper in the
World Water Week, 15–20 August, Stockholm, Sweden: 9 pp.
Poland, J.F. (1985). Guidebook to studies in land subsidence due to groundwater withdrawal. Studies
and Reports in Hydrology, 40. UNESCO, Paris, France: 350 pp.
Postel, S. (1999). The Pillar of Sand. W.W. Norton and Co., New York, USA: 313 pp.
Price, M. (2002). Who needs sustainability? In: K.M. Hiscock, M.O. Rivett & R.M. Davison (eds.),
Sustainable Groundwater Development. Balkema Publishers, Lisse, the Netherlands: 191–207.
Sahuquillo, A. & Lluria, M. (2003). Conjunctive use: a potential solution for stressed aquifers. Social
constraints. In: M.R. Llamas & E. Custodio (eds.), Intensive Use of Groundwater. Challenges and
Opportunities. Balkema Publishers, Lisse, the Netherlands: 157–177.
Selborne, J. (2000). The Ethics of Freshwater Use: A Survey. World Commission on the Ethics of Science
and Technology. UNESCO, Paris, France: 58 pp.
Shah, T. (in press). Groundwater and human development: challenges and opportunities in livelihoods
and the environment. In: Proceedings of the Stockholm World Water Week. Stockholm International
Water Institute, Sweden.
United Nations (2000). United Nations Millennium Declaration. A/RES/55/2.
United Nations (2003). Water for People, Water for Life. UNESCO-WWAP. Paris, France.
Vives, R. (2003). Economic and social profitability of water use for irrigation in Andalusia. Water
International, 28(3): 326–333.
WCD (World Commission on Dams) (2000). Dams and developments. A new frame for decision-making.
Earthscan: 404 pp.
CHAPTER 11

Is intensive use of groundwater a solution to


world’s water crisis?

A. Mukherji
Department of Geography, University of Cambridge, UK

ABSTRACT: This short chapter is a comment on the chapter titled “Significance of the Silent Revo-
lution of intensive groundwater use in world water policy”, by M.R. Llamas & P. Martínez-Santos (this
volume). The main argument of Llamas & Martínez-Santos is that intensive use of groundwater offers
an important window of opportunity for solving the so called world water crisis. In response to this,
my comment is divided into three parts. In the first section, I try to ascertain the magnitude as well as
the causes of water crisis in the world. In doing so, I particularly look at water poverty of nations. In
the second section, I critically analyze the postulates of Llamas & Martínez-Santos and comment on
how viable their recommendation of mitigating water crisis or meeting Millennium Development Goals
(MDG) with intensive groundwater use is likely to be. In the third, and the final section, I present some
of my thoughts on groundwater governance, which I believe will be vital, if groundwater is ever poised
to tackle world’s water problems.

Keywords: Groundwater, governance, water poverty, water crisis

1 WATER CRISIS: A MYTH OR A REALITY?

That the world is headed towards an impending water crisis is a cry heard with increasing
intensity in the recent years and that too from different quarters, be it pro-privatisation neo-
liberals or anti-globalisation groups, or be it advocates of large dams or opponents of it.
Though the fact that there will be water crisis in very near future is more or less agreed upon,
but there is disagreement about the precise nature of this crisis. Thus, one section of scholars
whom I call the neo-Malthusian pessimists have contended that there will be actual physical
scarcity of water because demand for water (due to population growth) will far outpace the
supply for water. However, this view stands challenged in the face of past evidence which
showed that human ingenuity in the face of scarcity is stupendous – more so in recent times
because such ingenuity is supported by advances in science and technology. An apt example
is the success of Green Revolution in India, which disproved prophesy of India’s doom quite
effectively. Again, there is enough evidence to show that as economies prosper, water use
efficiency goes up and the nature of demand changes. Thus Gleick (2000) rightly points out
that almost all water demand forecast done since 1970s have grossly over estimated water
demand based on historical rates, while in reality, water use efficiency has gone up, thereby
reducing water demand. The second group of scholars whom I called cautious optimists too
have voiced their concerns about future water crisis, but they contend that water crisis is rarely
a physical crisis, but more often a crisis of governance. Thus Seckler et al. (1998) predicted
182 A. Mukherji

that by 2025, demand for water will outpace that of supply by anything from 57% to 25%.
Cosgrove & Rijsberman (2000) made similar prognosis. Both suggested improving water
use efficiency and constructing additional supply sources as possible remedy of such water
crisis.
Again, the concept of water crisis, as simple it looks at the outset is in reality quite complex.
Various definitions of water scarcity are in vogue, such as physical water scarcity or demo-
graphic water scarcity like the Falkenmark’s index, IWMI’s (Seckler et al., 1998) concept of
absolute and economic water scarcity and Wolfe & Brooks (2003) concepts of first order,
second order and third order water scarcity. Similarly, measures to overcome water scarcity
are just as diverse, ranging from supply augmentation to demand management to a judicious
combination of both. One rather interesting work in this regard is that by Wolff & Gleick
(2002), where they talk of hard and soft path of water management.

1.1 Water poverty index and its determinants


A very recent development in quantifying water scarcity and its multidimensional aspects has
been the development of Water Poverty Index (WPI) by scientists at University of Keele and
Centre for Ecology and Hydrology, Wallingford, UK (Sullivan, 2002; Lawrence et al., 2002).
This index provides relative ranking of 147 countries based on a composite index comprising
of 5 sub indices, viz. water resource, access, capacity, use and environment. While WPI may
be criticized over the choice of variables and whether all variables deserve inclusion in the
first place and equal weight in the second place, yet it is a significant contribution because it
puts the multi dimensionality of water poverty squarely into picture. In addition, by focusing
on country wide situation, it presents a balanced view in that a country’s water crisis is not
exaggerated due to few basket cases, such as groundwater depletion in North Gujarat or arsenic
contamination in Bangladesh. Third and most important, this index provides some food for
thought in the form of understanding the relationship between physical water scarcity and
water poverty of nations and effectively sheds light on what determines water poverty of a
nation. Using simple graphs, I will make the following points:

(1) Water poverty of nation has very little to do with physical scarcity of water (Figure 1).
(2) And everything to do with Human Development Index (HDI) and per capita Gross National
Product (GNP) (Figures 2a, 2b).
(3) Demand for water (depicted by water use index in WPI) has little or no relation with
supply of water (depicted by resource index in the WPI) (Figure 3).
(4) On the other hand, water use index has very interesting relation with per capita GNP. As
per capita GNP goes up, water use increases till a certain threshold value and after that it
declines (Figure 4).
(5) Access to water, which is a component of WPI and indeed the true reflection of water
poverty of nations has again very little relation with water resource availability (Figure 5).
(6) However, access to water is strongly and logarithmically correlated with per capita GNP
(Figure 6).
(7) Finally water environment index, which is a sub index of WPI and is comprised of indices
of water quality, water stress, environmental management, information capacity and bio
diversity, is related to per capita GNP in an interesting way reminiscent of environmental
Kuznets curve (Figure 7).
Is intensive use of groundwater a solution to world’s water crisis? 183

120
WPI and Falkenmark Index

100
y = 0.104x + 49.032
80 R2 = 0.1804

60

40

20

0
1 7 13 19 25 31 37 43 49 55 61 67 73 79 85 91 97 103 109 115 121 127 133 139
Countries arranged in ascending order of Falkenmark Index

WPI Falkenmark ('000 m3 of water per capita per year) Linear (WPI)

Figure 1. Relationship between Water Poverty Index and Falkenmark’s index.

90
80
70
60
50
WPI

y = 6.4601 Ln(x) + 2.4287


40 R2 = 0.5558
30
20
10
0
0 5000 10000 15000 20000 25000 30000 35000 40000
GNP per capita in $PPP

Figure 2a. Relationship between Water Poverty Index and per capita Gross National Product ($PPP)1 .

Preliminary and unsophisticated as the above analysis is, yet it brings home several import-
ant points. First, neither demand for water, nor access to water has much to do with resource
endowments; instead both are strongly correlated to levels of income. While in case of access
to water (which in my view is the true indicator of water poverty or the lack of it) is posi-
tively and significantly related to per capita GNP, relationship of water demand with GNP
is a bit more complicated and is best captured by a slightly inverted U shape curve. This
means that as nations develop, water use efficiencies go up, a point corroborated by Gleick
(2000) with help of examples from the USA. Second, water environment is a quadratic func-
tion (U shape curve) of per capita GNP, signifying that in the initial stages, nations develop

1
PPP means purchasing power parity. This is generally used to ensure comparability across different
currencies.
184 A. Mukherji

90
80
70
60
50
WPI

y = 44.178x + 27.02
40
R2 = 0.654
30
20
10
0
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
HDI

Figure 2b. Relationship between Water Poverty Index and Human Development Index.

18
16
14
Water use (demand)

12
10
8
6
4 y = -0.0076x + 9.9305
2 R2 = 0.0002
0
0 5 10 15 20 25
Water resource (supply)

Figure 3. Relationship between water demand and water supply.

18
16
14
Water use index

12
10
8
6
4
y = -7E-09x2 + 0.0001x + 10.023
2 R2 = 0.1448
0
0 5000 10000 15000 20000 25000 30000 35000 40000
GNP per capita in $PPP

Figure 4. Relationship between water demand and per capita Gross National Product.
Is intensive use of groundwater a solution to world’s water crisis? 185

25

20

15
Access

10
y = 0.0695x + 12.568
R2 = 0.0029
5

0
0 5 10 15 20 25
Resources

Figure 5. Relationship between access to water and water resource availability.

25

20
Access to water

15
y = 3.706 Ln(x) - 17.988
R2 = 0.685
10

0
0 5000 10000 15000 20000 25000 30000 35000 40000
GNP per capita in $PPP terms

Figure 6. Relationship between access to water and per capita Gross National Product.

40000

35000 y = 492.5x2 - 8797.4x + 43898


R2 = 0.402
30000

25000
GNP

20000

15000

10000

5000

0
5 7 9 11 13 15 17 19
Environment

Figure 7. Relationship between water environment and Gross National Product.


186 A. Mukherji

Table 1. Determinants of water poverty of nations.

Standardized beta coefficient


Dependent
variable Intercept Falkenmark index HDI GNP per capita in $PPP terms R2

WPI 27.119* 0.262* 0.781* – 0.722


WPI 50.002* 0.282* – 0.606* 0.482

* Denotes significance at .001 level for two tailed t test.

at the cost of environment, but as development crosses a desirable threshold, demand for
environmental goods and services increase and so does environmental quality. This is very
similar to Kuznets environmental curve. Finally, water scarcity (or water poverty) is rarely a
function of overall water resource availability in the country, but is more a function of level
of economic and social development as reflected by HDI and GNP. In other words, water
poverty simply reflects economic underdevelopment. The following two regression equations
depict this.
In sum, therefore, physical scarcity of water has very little to do with water poverty of
nations, while wealth generating capacity of a nation has. In effect, as economies prosper, water
use efficiencies go up and this has much to do with availability of better technologies, which
developed nations can afford. But at the same time, any effort to transpose such efficiency
enhancing technology to a not so developed country is unlikely to succeed. Thus, in the long
run, economic and human development is the best way to alleviate water poverty.

2 CAN INTENSIVE GROUNDWATER USE SOLVE WORLD’S WATER PROBLEMS?

Although, physical water scarcity is rarely a problem, yet it is undeniable that there exist severe
problems regarding access to water. Thus, millions of people in the developing world do not
have access to safe drinking water and sanitation – a number which United Nations MDG aims
to half by the year 2015. In this context, Llamas & Martínez-Santos (this volume) contend
that intensive and responsible use of groundwater has the potential of solving much of world
water problems.
In support of their views, they present the following advantages of groundwater over surface
water. First, using examples from India and Spain, they show that technical efficiency or water
productivity (m3 /ha) of groundwater is several times higher than that of surface water. Second,
in addition to technical efficiency, both economic ($/m3 ) and social (jobs/m3 ) productivity
of groundwater is higher than any other source of irrigation. Third, unlike centralized and
bureaucratic development of most surface water systems, development of groundwater has
been dispersed and democratic (also see Deb Roy & Shah, 2003; Shah et al., 2003; Mukherji &
Shah, 2005a, 2005b). This is what the authors appropriately call the silent revolution. In add-
ition, unlike surface water schemes, groundwater extraction is rarely subsidized, so in effect
farmers pay the full cost of groundwater extraction. Even in India, where electricity is highly
subsidized, farmers using groundwater still pay anything between 3–5 times the price for water
than do farmers in the canal command area. The fact that each unit of groundwater extraction
entails additional cost, ensures that groundwater is used sparingly, resulting in higher water
use efficiency as mentioned earlier. Since groundwater irrigation is under individual control,
Is intensive use of groundwater a solution to world’s water crisis? 187

use of complimentary resources (HighYieldingVarieties – HYV – seeds and fertilizers) is high,


which again results in higher crop productivity (Shah, 1993; Kahnert & Levine, 1993). The
authors also point out that groundwater irrigation involves less corruption, first because costs
are much lower than even a small scale surface water scheme, and second because, in most
cases, groundwater structures have been self-financed by the farmers. Though this is largely
true, there have been allegations of corruption even in groundwater sector in India, especially
when it involved any kind of subsidy scheme (Pant, 2003), but it is not nearly anywhere of
those reported from surface water bureaucracy in India (Wade, 1984). Due to inherent advan-
tages of groundwater use, it has important poverty alleviation implications in water abundant
regions, a point noted by the World Bank as early as in 1989 (Kahnert & Levine, 1993). Quite
often, a rather unexpected effect of intensive groundwater use has been the political empower-
ment of farmers. A case in point is the recent electoral upset in India, where it is claimed that
farmers voted out a government which was doing exceedingly well in macro economic front
(high growth in Gross Domestic Product, historically high levels of foreign exchange reserve,
high foreign institutional investment and so on), but had failed to take into account farmers’
demand in the form of lower input prices and higher output prices. An analysis of states where
existing government was toppled correlates very well with the level of groundwater use. Thus,
the agriculturally developed states like Punjab, Haryana, Tamil Nadu, Andhra Pradesh and
Karnataka have all voted for a change in government and all these states are high users of
groundwater. In fact, the issue of electricity reforms (mostly financed by either World Bank
or Asian Development Bank and which entailed removal of electricity subsidy for agriculture)
has been at the forefront of farmer’s unrest in these states. So much so, the new Chief Minister
of Andhra Pradesh has promised free electricity to the farmers on the very day he took oath of
office, and the neighboring Chief Minister of Tamil Nadu has followed suit. Certainly these
populist policies have deleterious effect, but the point I am making is that farmers who have
enjoyed fruits of development (mostly due to groundwater led Green Revolution) also become
politically vocal. But in other parts of India, such as eastern states of Bihar and West Bengal
where level of agricultural development is still low (as is level of groundwater use), farmers
lobby is either weak or is almost non-existence.
Llamas & Martínez-Santos (this volume) also point out certain misconceptions in vogue
about groundwater use, which they refer to as hydromyths. Two such important hydromyths are
that groundwater is a fragile resource and as such should be used as a last resort; and second,
that rate of extraction of groundwater should be never more than rate of recharge. They go on to
show that both these conceptions are erroneous and at best lead to confusion and apathy about
the benefits inherent in groundwater use. On the whole, authors are optimistic and enthusiastic
about the positive effect of intensive groundwater use. But to be fair, they also point out certain
genuine drawbacks of groundwater, which stems mostly from mismanagement and hence can
be corrected through improved management practices. However, this is easier said than done.
They also note that the concept of overexploitation of groundwater often lacks scientific
rigour and that there is rarely any aquifer in the world (above 1000 km2 ) that has shown
significant levels of dewatering. Perhaps true. This also implies that at a broad level, there is
no groundwater resource crisis per se and whatever crisis remains can be adequately managed.
But, although true at a macro level, this argument breaks down at the level of a farmer. For
example, in the hard rock aquifers in India, farmers have been forced to commit suicide due to
crop failure, which is very often, though not always related to well failure. It might very well
be that the aquifer is not really dewatered as is often claimed in popular newspapers, but the
188 A. Mukherji

fact remains that at the given level of economic status of the farmers, they could not deepen
their wells further, neither could they grow crops which fetched them higher market value.
In this circumstance, debating the exact definition of aquifer overexploitation will but seem
irrelevant to affected farmers, though such definitional debate remains relevant for the water
managers. Thus, the concept of aquifer overexploitation has as much social and economic
implication as it has technical meaning. Having extolled the virtues of groundwater use and
to a limited extent the possible negative externalities, the authors lay emphasis on awareness
creation and stakeholder participation for fostering sustainable groundwater management.
This paper is a sterling contribution for various reasons. For one, it rightly stresses the
positive benefits of groundwater, which is rarely emphasized in mainstream water resources
discussions. Thus, while Cosgrove & Rijsberman (2000) and Seckler et al. (1998) discuss water
future 2025, they do not pay any attention to the fact that today more than 50% of world’s
irrigated area is under groundwater irrigation and that water use efficiency in groundwater
irrigated areas is anytime between 2 to 5 times more than surface water irrigation! Second,
this paper is a welcome change in a sea of pessimistic doomsday sayers such as one that
claims that almost 10% of world’s food production is at the mercy of groundwater overdraft
to the tune of around 200 km3 and that such pillar of sand as it were is bound to collapse
sooner or later (Postel, 1999). Third, the authors’ attitude towards environmental value of
groundwater is pragmatic, quite in contrast to the rather dogmatic views often imposed by
western scholars on the developing world. They rightly comprehend the fact that utopian ideal
of preserving groundwater for the future generations and for nature without providing any
options for the present generation in poor countries is unlikely to succeed. They also effectively
highlight the changing attitude to environmental values as society’s make economic progress
by citing examples from England, Spain and Turkey. Finally, by emphasizing such aspects as
information sharing and stakeholder participation, they tend to demystify the hard and often
incomprehensible science of hydrology for the benefit of common people.
However, the chief drawback of the chapter lies in its failure to take into account the negative
aspects of groundwater use more thoroughly. In fact, negative externalities are more or less
brushed aside in their optimistic everything is all right with groundwater attitude which is in
stark contrast with the dominant pessimistic paradigm of groundwater and its ill effects. I claim
that truth lies somewhere in between pessimistic doomsday saying and optimistic every thing’s
all right attitude. It is true that groundwater economy is indeed booming and has conferred
huge socio-economic benefits in large parts of the developing world. But equally true is the
fact that this boom is likely to burst without proactive governance and when it does, the ill
effects will nullify or even surpass the social gains made so far (Mukherji & Shah, 2005a).
This is precisely the issue I will deal with in the next section.

3 GROUNDWATER GOVERNANCE

In this section, I will lay down the conditions under which intensive use of groundwater is
likely to confer benefits with minimum negative externality. To do this, I will claim that
policies have to shift from development to management and finally to governance mode.
The present water crisis is “mainly a crisis of governance” – so declared the Global Water
Partnership (GWP, 2000). This crisis of governance in the water sector is no where as visible
as it is in groundwater. Figure 8 diagrammatically portrays the overall concept of governance.
Is intensive use of groundwater a solution to world’s water crisis? 189

The concept of governance has found acceptance among water professionals and GWP (2000)
has defined water governance “as a range of political, social, economic and administrative
systems that are in place to regulate the development and management of water resources and
provision of water services at different levels of society”. However, current discourse on water
governance has not paid any special attention to groundwater per se, assuming that the essen-
tial elements of governance remain same across various water sectors, which is perhaps true
at a certain conceptual level. However, equally true is the fact that governing groundwater is a
greater challenge because: firstly, it is an invisible and fragile resource; and secondly, because
intensive groundwater use of the scale we witness today is a recent phenomenon (of the last 30–
50 years or so), many of the challenges encountered are new, and at times unexpected. A very
good example of the unexpected nature of groundwater problem is the catastrophic arsenic poi-
soning in Bangladesh and West Bengal. Here, in order to provide safe and pure drinking water,
World Bank and other donors encouraged extraction of groundwater through shallow tubewells
little knowing that they were substituting the fear of surface water borne diseases with one of
massive mass poisoning episode in human history (Centre for Science and Environment, 2002).
The first step in groundwater governance is adequate and good quality information, not only
hydrogeological, but also socio-economic. In many instances, such information is missing or
more possibly not accessible in the public domain due to unwillingness of the groundwater
experts to share it with general public. While hydrological data is systematically gathered, if not
disseminated, collection of social and economic data is few and far between. India undertakes
reasonably detailed Minor Irrigation Censuses every five year, but results are published with
great delay. Access to good quality data is perhaps less of an issue in developed nations,
yet hardly any country so far has 50 years or so timeline hydrological data that some like
Llamas (pers. comm.) contend is needed for thorough understanding of complex hydrological
phenomena.
Secondly, what ails groundwater sector, very often, is the misallocation of roles and
responsibilities assigned to various organizations that are in the business of managing ground-
water. Thus, central and provincial governments are very often given responsibilities that are
beyond their human and financial resources, while farmers and other stakeholders are asked

Global

Regional and
sub regional

Civil society Nation state Market

Local self
governments

Figure 8. The new and idealized concept of governance.


190 A. Mukherji

Global
lessons

Regional and
Civil society sub regional
Markets

Central governments

Local self
governments

Present role in Desired role in Desired level of


groundwater governance groundwater governance dynamic interaction
(size of circles denote in future (size of circles in future (denoted by
importance). denote importance). respective widths).

Figure 9. The existing and the desirable groundwater governance scenarios.

to participate in aquifer management when their direct interests lie in non-participation. The
problems inherent in groundwater governance are denoted in Figure 9. It shows the present
state of governance and the desired governance structure and process that will possibly impart
a modicum of order in anarchic groundwater economy. Developments in groundwater sector
are guided very little by government but are a result of an interplay between many actors and
factors in which state and government has much less central and dominant position compared
to, say, in the surface water sector. But at the policy circles, there is very little recognization of
the growing contribution of groundwater led economy, that too an economy that is developed
and sustained by millions of small and marginal farmers. Thus, better groundwater govern-
ance it will mean recognizing greater role for the markets, the civil society and the local self
governments and much less role for the central and provincial governments.
Finally, there is a need for attitude shift among the water policy makers and practition-
ers. Thankfully, in the field of groundwater, it has now come to be widely recognized that
Is intensive use of groundwater a solution to world’s water crisis? 191

Groundwater economics,
institutions and society

Groundwater hydrogeology; Groundwater


Groundwater flow modeling; policy and
Salinity balance studies; governance
Groundwater pollution;
Artificial recharge techniques
and models;
Groundwater estimation

Present state of knowledge Desired state of knowledge

Figure 10. The knowledge development challenge of groundwater governance.

management is not the exclusive forte of hydrogeologists. In fact, in understanding the chal-
lenge of sustainable groundwater management anywhere in the world, it is critical to blend three
distinct perspectives: (a) the characteristics and behavior of groundwater resource (resource
perspective); (b) the characteristics and behavior of resource user communities (user perspec-
tive); and (c) the institutional frameworks under which the resource is appropriated and used
(institutional perspective) (see Figure 10).
So far, global knowledge development as well as capacity building on groundwater use
in agriculture has been dominated by the resource perspective; and a critical value-adding
contribution is to be made in expanding global knowledge and capacity in user and insti-
tutional perspectives. Undergraduate and Masters’ courses on hydrogeology in South Asia,
for instance, spend less than 10% of teaching time on economics and sociology. Possibly
same holds true in developed countries. When these students grow up as researchers and pol-
icy makers in groundwater sectors, their worldview tends naturally to get dominated by the
resource perspective. And since economists and sociologists seldom foray into groundwater
research, user and institutional perspectives have remained largely unexplored in the existing
groundwater knowledge base.

4 CONCLUSIONS

From the foregoing sections, we saw that water crisis is very rarely a crisis of physical avail-
ability of water per se; it is much more a crisis of governance. More importantly, water crisis
is but a symptom of the larger malaise of poverty that still plagues this world in this 21st
192 A. Mukherji

century. Therefore, it follows the best way to tackle water crisis is to tackle poverty. However,
it almost seems as if there is a global conspiracy aimed at keeping the poor people poor in per-
petuity. This is manifested in such concepts as sustainable use of resource, which for instance
talks of inter-generational equity, while forgetting that there exists a deep intra-generational
gap – the gap between the rich and the poor countries. Second, associated with the concept
of sustainability is the concept of environmental protectionism which often in the name of
environment stalls development. But empirical evidence shows that in the initial years of
development, environment takes a backseat, while as soon as economies prosper and demand
for environmental goods and services increase, quality of environment too improves. Third,
very often there is an attempt to impose techniques (e.g. of increasing water efficiency) that
work well in the developed nations, to the poor countries with deleterious effect. Depend-
ing on the social and economic status of every nation, its priorities vary and rightly so. Any
attempt to tinker with this is a formula for disaster, as has been seen in case of piped domestic
water supply in most African countries. Finally, to deny development to poor countries in the
name of environment is but a ploy to maintain the status-quo of inequitable distribution of
wealth.
In this context of water crisis, which is but a symptom of poverty, role of groundwater is
commendable. For one, it is relatively cheap to develop groundwater for irrigation; second,
its development is decentralized and lies in the hands of the farmers; and third, because of
inherent features of groundwater as a resource, it tends to be pro-poor. Thus, groundwater
use and benefits thereof have deeply impacted humankind and it will not be exaggeration to
say that groundwater use in agriculture has indeed revolutionized food production and created
livelihood opportunity for millions of people. But, like every other natural resource, there are
both costs and benefits and the prime aim in governing groundwater is to ensure that costs
(or negative impacts) of intensive groundwater use does not exceed the benefits thereof. The
costs or negative externalities hit the poor hardest and hence a word of caution is in place.
Governing groundwater is a challenge, and if properly undertaken, it does have the potential
of denting poverty, as it has already done in large parts of South Asia.

REFERENCES

Centre for Science and Environment (2002). Down to Earth, Volume 11(1), May 31: 38.
Cosgrove, W.J. & Rijsberman, F.R. (2000). World Water Vision: Making Water Everybody’s Business.
Earthscan Publications, London, UK.
Deb Roy, A. & Shah, T. (2003). Socio-ecology of groundwater irrigation in India. In: M.R. Llamas &
E. Custodio (eds.), Intensive Use of Groundwater. Challenges and opportunities. Balkema Publishers.
Lisse, the Netherlands: 307–336.
Gleick, P.H. (2000). The changing water paradigm: a look at twenty-first century water resources
development. Water International, 25(1), March: 127–138.
Global Water Partnership (2000). Towards Water Security: a Framework for Action. March.
Kahnert, F. & Levine, G. (eds.) (1993). Groundwater irrigation and rural poor: options for development
in the Gangetic basin. World Bank, Washington, D.C., USA.
Lawrence, P., Meigh, J. & Sullivan, C. (2002). The Water Poverty Index: an international comparison.
Keele Economics Research Papers, 2002/19. Keele University, UK: 24 pp.
Mukherji, A. & Shah, T. (2005a). Groundwater socio-ecology and governance: a review of institutions
and policies in selected countries. Hydrogeology Journal, 13(1): 328–345.
Mukherji, A. & Shah, T. (2005b). Socio-ecology of groundwater irrigation in Asia: an overview of issues
and evidence. In: A. Sahuquillo, J. Capilla, L. Martínez Cortina & X. Sánchez Vila (eds.), Selected
Is intensive use of groundwater a solution to world’s water crisis? 193

Papers of the Symposium on Intensive Groundwater Use (SINEX). IAH Selected Papers. Taylor and
Francis Publishers, the Netherlands.
Pant, N. (2003). Key trends in groundwater irrigation in the eastern and western regions of Uttar Pradesh.
Paper submitted to IWMI-Tata Water Policy Program, Anand, Gujarat, India.
Postel, S. (1999). The pillar of sand. W.W. Norton & Co., New York, USA: 313 pp.
Seckler, D., Amarasinghe, U., Molden, D., de Silva, R. & Barker, R. (1998). World water demand
and supply, 1990–2025: scenarios and issues. IWMI Research Report No. 19, International Water
Management Institute, Colombo, Sri Lanka.
Shah, T. (1993). Groundwater markets and irrigation development: political economy and practical
policy. Oxford University Press. Bombay, India.
Shah, T., Deb Roy, A., Qureshi, A.S. & Wang, J. (2003). Sustaining Asia’s groundwater boom: an
overview of issues and evidence. Natural Resources Forum, 27: 130–141.
Sullivan, C. (2002). Calculating a world poverty index. World Development, 30(7): 1195–1210.
Wade, R. (1984). The market for public office: why the Indian state is not better at development.
Discussion paper 194, Institute of Development Studies, University of Sussex, Brighton, UK.
Wolfe, S. & Brooks, D.B. (2003). Water scarcity: an alternative view and its implications for policy and
capacity building. Natural Resources Forum, 27(2), May: 99–107.
Wolff, G. & Gleick, P.H. (2002). The soft path for water. In: P.H. Gleick, World’s Water 2002, Chapter 1.
VI

Water and Poverty Alleviation


CHAPTER 12

Water and poverty alleviation: the role of investments


and policy interventions

R. Bhatia
Resources and Environment Group, New Delhi, India

M. Bhatia
Department of Agricultural and Resource Economics, University of California, Davis, USA

ABSTRACT: This chapter shows that investments in water infrastructure (irrigation and hydropower
dams, small check dams) provide direct and indirect benefits to the rural poor and agricultural labor that
are higher than those obtained by large farmers and urban households. Improved management of water
resources, through flexible allocation among sectors and users, provides benefits to the poor and can be
an important instrument of poverty alleviation in developing countries. There is a need for monitoring
the impact on the poor of community-managed rural water supply schemes where full cost recovery from
the users is mandatory. Scaling-up of islands of success in community participation in urban sanitation
faces constraints such as sources of public funds, cost recovery and financial viability, leveraging of
community funds and private financial resources. To ensure protection, augmentation and equitable
distribution of water resources, innovations will be required in water rights, pricing of water, targeted
and cross-subsidies.

Keywords: Impact of dams, flexible water management, community participation in water and
sanitation, water and poverty, water rights, targeted subsidies

1 INTRODUCTION

This chapter presents an overview of the role of policy interventions, investments in water
infrastructure and improved management of water resources and services in poverty alleviation
in developing countries. A number of case studies have been presented to show how water
resources development and management can affect the rural and urban poor, both directly and
indirectly, by providing livelihoods, food security, lower food prices and higher income and
wages for all. Section 2 presents an overview of various dimensions of poverty and the Water
Poverty Index (WPI) in the context of the sustainable livelihoods framework. Section 3 provides
an overview of policy, investments and management interventions required in the water and
related sectors for poverty alleviation/reduction. Section 4 presents empirical evidence to show
that investments in irrigation and hydropower projects have considerable multiplier (direct and
indirect) benefits to farmers and agricultural workers in terms of increased employment, real
wages and income in agro-industries.
Section 5 shows how improved management of water resources can benefit the poor. Sec-
tion 6 presents a review of experiences in community management of rural water supply
198 R. Bhatia & M. Bhatia

services and community participation in urban sanitation in the context of the Millennium
Development Goals of providing water and sanitation to millions of poor people in Asia and
Africa. The chapter provides a summary of constraints to the scaling-up of these islands of suc-
cess. In section 7, major policy issues are analyzed that affect investments in water resources
development, protection, management and provision of services to the poor. These include
legislation and regulation, water rights, pricing of water and financial sustainability of water
investments, cross subsidies, and irrigation pricing in the context of subsidies provided by
developed countries to their farmers. The last section presents some conclusions.

2 WATER AND POVERTY

2.1 Dimensions of poverty


Although politicians have always used poverty removal or reduction as a slogan, millions of
people continue to live below the poverty line in Asia and Africa. The living condition of large
populations, particularly in the slums, is a matter of shame for all those who enjoy the benefits
of a higher standard of living. According to the State of World Rural Poverty produced by the
International Fund for Agricultural Development (IFAD) in 1992, out of a population of some
4000 million people living in 114 developing countries, more than 2500 million live in rural
areas, and of these approximately 1000 million live below the poverty line. More than half
of this population lives in highly degraded lands. These people suffer from a lack of basic
necessities like safe drinking water, adequate food and health care, which means that almost a
third of the people in the developing world have a life expectancy of just 40 years. The IFAD
report says that less than half the rural population had access to safe drinking water and even
less to irrigation water to ensure sustained agricultural production.
It must be kept in mind that these indicators would vary from one country to another
and from one social group to another and over time. For example, in the case of rural water
supply in India, a community is classified as having safe drinking water supply if a source
(government-owned) is available within a given distance (say 100 m) from the dwellings.
Using this approach, it is claimed that 80% of rural households have access to safe drinking
water. Such an approach is misleading because the indicator includes only public sources of
supply (excludes private sources, which account for a high percentage in some states such as
Punjab and Kerala). The method also does not take into account whether the source of water
supply is functioning or not, whether water quantity is adequate for all the households (as
per norms of 50 L/d per capita), and whether the water quality is good or bad (contaminated,
saline or with arsenic or fluoride). Such measurements will have to be improved if these have
to be used as indicators of meeting basic needs in the classification of poverty of regions or
households.

2.1.1 Poverty and resource use


As may be seen from any empirical estimates of poverty levels, the poor comprise heteroge-
neous groups in conflict with each other for resources, particularly for land, water and energy.
There are a number of studies (Chopra et al., 1990; Meinzen-Dick et al., 1997a, 1997b; Bos &
Bergkamp, 2001; Meinzen-Dick & Bakker, 2001; Soussan, 2001), that show that competition
among the poor for these, usually common property resources, leads to overexploitation and
degradation of resources and conflicts among users. Such degradation of resources results
Water and poverty alleviation: the role of investments and policy interventions 199

in reducing the ability of these resources (e.g. watersheds, wetlands, tanks, lakes, rivers)
to provide livelihoods to a large number of poor people who do not own land or water
resources.
Most development agencies and government institutions concerned with the imple-
mentation of poverty reduction programs, either from a supply-dominated approach or a
demand-based approach, has to start by dealing with the issues related to quality and quantity
of water resources. It is necessary to break the vicious circle that is created among the poor,
who need water and other resources for their survival and livelihoods and overexploit these in
their desperation.
Thus, starting from the supply-dominated approach we must distinguish between water
service delivery problems among the rural poor and the urban poor. In many countries, it is
the urban poor who face the serious health consequences of inadequate water supply, polluted
water sources/supplies and high costs of provisioning and coping strategies. In many countries
in Asia, urban populations are rising very fast and it is expected that almost 50% of total
population in South Asia will live in urban areas. Majority of the people in Latin America
are already in urban areas. Meeting the basic needs of water and sanitation services of urban
poor at affordable prices is going to be a major challenge for the next two decades. The
capital costs of providing basic services in congested urban areas are so high that innovative
solutions will have to be found so that adequate quantities of safe water are available to the
urban poor.

2.1.2 Vulnerability of the poor


The bulk of the world’s poorest people, 800 million to 1000 million, live in arid areas and
depend directly on natural resources including water for their livelihoods (UN, 1999, 2004).
An important dimension of water and poverty is the vulnerability of the poor people to shocks
or unexpected changes in water regimes (e.g. droughts and floods) and their consequences
(e.g. decline in market prices of their outputs or rise in food prices). Millions of people, mainly
the poor and disadvantaged, remain at risk from the lack of clean water, public health risks
from inadequate sanitation affect some 50% of the world’s population, and the number of
people at risk from floods and drought continues to rise. At the same time risks from degraded
ecosystems have increased inexorably, wetlands have been destroyed, over abstraction has
lowered water tables and caused major rivers to cease flowing to the sea, and both ground and
surface waters have been grossly polluted (Rees, 2001).

2.1.3 Feminization of poverty


Another important aspect of poverty relates to the feminization of poverty and its effects
related to water and other resources. Available literature (UN, 2004) shows how poverty is
related to ownership of property rights, especially land rights. From studies in India, it has
been shown that the single most important economic factor affecting women’s situation is the
gender gap in command over property, an issue virtually neglected in research, policy and
grassroots action. Based on a review of laws governing property and its inheritance, it has
been pointed out that women’s access to, and control over, landed property is crucial to their
empowerment (Agarwal, 1997).
On a more basic subsistence level, restricting land ownership puts women at a disadvantage
in providing for their families, gaining economic independence and enjoying personal security.
It limits their production possibilities, such as cultivating crops, trees or small gardens and
200 R. Bhatia & M. Bhatia

keeping livestock. Without these means to provide for themselves, women without land, as
well as their children, have a significantly higher risk of absolute poverty. In virtually every
legal system, gender inequality in land issues remains. For instance, some systems allow only
small plots of land for women; others restrict the conditions under which women can inherit
land and some are age discriminatory, prohibiting access to land for young women or the
elderly. Women’s empowerment in environmental decision-making and land management is
well documented as crucial for sustainable human development. In recognition of this, the
United Nations Development Program (UNDP), the United Nations Development Fund for
Women (UNIFEM) and other UN agencies are working to help countries transform their legal
and social structures and prioritize land rights for women1 .

2.1.4 Ecological poverty


According to Agarwal (1999), most rural people in the developing world live in an economy
built on natural capital, which can also be described as a biomass-based subsistence economy.
Ecological poverty is invariably the main cause of their impoverishment and an inability to
meet their basic survival needs. Ecological poverty is defined as the lack of natural resources,
both in quantity and quality, that are needed to sustain a productive and sustainable biomass-
based economy. An excellent indicator of ecological poverty is the amount of time a rural
household spends on collecting basic survival needs like water, fuel and fodder (for animals)
from the local environment.
In conditions of high ecological poverty, this time can be excessive leading to enormous
work pressures on the household. Since in most cultures, these activities are carried out by
women, it is they who have to bear the maximum impact of ecological poverty. As pointed
out by Agarwal & Narain (1980), degraded drylands and mountain sides can pose the biggest
challenge for reducing female work burden. In a study carried out by the Center for Science
and Environment (CSE), New Delhi, in a village of 213 people, the total work-hours spent in
animal care, agriculture, fuel and fodder collection, animal grazing and other household and
market-related activities during an entire year was 366,156, of which women contributed 58%
of the labor and children another 26%. Men contributed a mere 15%. All this had a major
impact on female literacy. Fortunately, the village had a school of its own. Most children
went to school late but by about age 10 almost all children in the village were in school. But
for girls education was very brief. No girl in the age group 15–20 was in school whereas
78% of the boys in this age group were still receiving education. No girl in the village had
reached the high school level. In fact, less than 10% of the girls went beyond the primary
stage.
As a contrast to the situation found in many states in India, women in Kerala (India)
rarely spend more than an hour collecting basic needs like fuel, fodder and water and their
work burden is far less than that of women in other parts of India. Kerala has often been
highly praised by demographers for its low birth rate and high female literacy despite its
low per capita incomes. But it may be useful to investigate whether the high availability of
biomass and water in the tropical, humid environment of Kerala – the state is blessed with

1
The Secretary General’s report on improvement in the situation of women in rural areas (document
A/52/326), underlines the critical importance of reconciling and strengthening the productive and
reproductive roles of women farmers and entrepreneurs in improving their situation.
Water and poverty alleviation: the role of investments and policy interventions 201

two monsoons a year – was an important factor in the success of the state’s female literacy
programme.
In rural areas where marginal farmers and landless laborers constitute 60% to 70% of
population (e.g. India), a large majority of poor people do not have access to land and water
for meeting their basic needs for survival and livelihoods. These poor groups then depend
on government/community sources for meeting basic needs. As a result, a large number of
women and girls spend a substantial part of their time in collecting water, fuel, fodder and
food. According to Agarwal (1995, 1997), in India, the availability of natural resources to
a large proportion of the poor rural population has been severely eroded over the past 20
years. The adverse class-gender effects include an increase in the time and energy that women
and girls spend in fuel, fodder and water collection, a decrease in women’s incomes from
non-timber forest products and agriculture, an adverse effect on their health and nutrition,
an erosion of women’s social support networks, and a decline in their traditional knowledge
of plants and species. The author maintains that the gender specificity of these effects stems
from pre-existing gender inequalities in the division of labor, the intra-household distribution
of subsistence resources, access to productive resources, other assets and income-earning
opportunities, and participation in public decision-making forums. According to Agarwal
(1995) rural women are worst off in regions where these three dimensions of disadvantage are
strong and reinforce each other, as in many parts of Northern India, and especially Bihar. They
are best off where all three types of disadvantages are weak, as in Southern and Northeast
India, and especially Kerala.

2.2 Water and livelihoods for the poor


Water resources availability (rainfall and irrigation) affects the livelihoods of poor people
by providing them with food (collected from community lands, forests, wetlands, tanks,
fish ponds), employment and wages (both in kind and cash). Livelihood thinking, which
developed in the 1980s as an alternative to production thinking, challenged beliefs about
the neutrality of technology and the absolute ability of experts to promote optimal production
systems. It also required a new professionalism to make resource management and technology
serve small farmers. Livelihood thinking involves understanding water environments and
technologies, understanding and working with the political processes through which local
groups can question water assessment and allocation mechanisms, including expert solutions,
and working directly with small farmers. This shift in orientation can foster local water-control
initiatives that support users in negotiating their rights to water and livelihoods – within both
water-basin and local water systems.
The dependence of poor people, particularly women, on common property resources, results
in the overexploitation and degradations of land, water and other natural resources. Poverty
shortens time horizons of the poor people leading them to overexploit resources. This means
that poor people basic needs and livelihood considerations lead them to prefer consumption
now over consumption in the future. The problem of Tragedy of the Commons gets exacerbated
for a watershed or common water body (wetland, lake) when poor people have to eke out a liveli-
hood from the limited natural resources and meet their basic needs of fuel, fodder and water.
The situation worsens for the poor when their hand pumps and wells become dry because rich
farmers pump water for irrigation, where private benefit of extracting groundwater becomes
more important than preserving a groundwater reservoir (belonging to a community) and
202 R. Bhatia & M. Bhatia

using it in a sustainable manner. This leads to a killing of the resource before its time. There
comes a time when the resource is gone (or groundwater table has declined) and is not there
for anybody’s use.
It may be emphasized that the poor comprise a heterogeneous group that may be in con-
flict with each other for the access to natural resources, such as land, water and energy;
so it is important to have a broad social participation to find the proper solutions. Further,
there has to be a growing awareness of the need to shift the design of programs and projects
targeted to reduce poverty (by increasing the water supply) into projects that have greater
emphasis on the vital connection that exists between land use, secure water accessibility and
poverty. Until now the health impacts of improvement of service delivery – either the supply-
dominated approach or the demand-based approach – were seen as important. However, it
has now become increasingly evident that the quality and quantity of water is a fundamental
requirement in order to escape the vicious circle that is created among the poor, who need
water and other resources for their survival and livelihoods and overexploit these in their
desperation.

2.3 The Water Poverty Index (WPI)


The Water Poverty Index (WPI) (Sullivan & Meigh, 2003; Sullivan et al., 2003) is an inter-
disciplinary tool that integrates the key issues relating to water resources, combining physical,
social, economic and environmental information associated with people’s ability to get access
to water and to use water for productive purposes. It is most relevant at the community or
sub-basin scales. The WPI can best be applied in practice to generate useful data that may be
used to generate benefits, especially for poor people who suffer from inadequate access to
water. WPI values would need to be generated over wide areas, and this would require sub-
stantial institutional development. This may require the use of existing census procedures and
simplified data collection, including widespread data collection through schools. A number
of technical issues relating to implementation of the WPI are:
(1) How the different spatial scales inter-relate; (2) how the assessment of the physical
resource and the collection of social and economic data may be made compatible; and (3) how
the WPI value can be used in practice, and what are the issues and problems that this presents.
The WPI concept is closely linked to the ideas of sustainable livelihoods and WPI is a way
of measuring water status focusing on poverty and the livelihoods components of the poor.
The five key components of WPI – Resources, Access, Capacity, Use and Environment – are
closely analogous to the livelihood capitals (infrastructure, physical capital, social capital, and
financial capital for investment. There is no one to one equivalence, but the WPI and sustainable
livelihoods concepts fit together (Sullivan et al., 2003: 193). In WPI, it is considered that a
lack of water is consistent with a lack of one of the basic prerequisites to an effective life,
but lack of water will have many additional repercussions. For example, low quantities of
water can be shown to have a direct relation to health, as personal and food hygiene will be
less effectively carried out. Further more, there are a number of illnesses that can result from
poor water quality or contaminated water. In terms of productivity, water is usually a factor,
even in subsistence households. For many poor households, their incomes from livestock,
fishing, critically depends on availability of water in canals, ponds, tanks and wetlands. Thus,
inadequate access to it (water) will impact on economic performance, and – obvious but
often forgotten – time spent collecting water will not be available for other activities (Sullivan
Water and poverty alleviation: the role of investments and policy interventions 203

et al., 2003). To redress any kind of poverty, access to water and other natural resources (along
with availability of the livelihood capitals) must be redistributed more equitably.

3 POLICY, INVESTMENT AND MANAGEMENT INTERVENTIONS FOR POVERTY


ALLEVIATION/REDUCTION

To meet the objective of poverty alleviation/reduction, a number of actions will be required


at the policy and institutional levels (including legal and regulatory issues) in the water sector
and related sectors. Substantial investments will be required in water infrastructure such as
check dams, hydro dams, multipurpose dams, canal systems and groundwater pumping. New
paradigms in management of water resources will be required that will ensure flexibility in
allocation of water among sectors, users and uses. Community involvement in the manage-
ment of water resources and in the provision of water and sanitation services can address the
problems of the poor and slum dwellers.
It may be pointed out that the poor benefit from a number of interventions, both directly
and indirectly. Some of these interventions, discussed in detail in the next sections, are:
– Investments in water infrastructure (e.g. check dams, multi-purpose dams, groundwater
projects) provide direct and indirect benefits to the poor, inter alia, in terms of higher
employment, higher real wages, increased non-farm employment and lower food prices.
– Improved management of resources (e.g. flexibility in sectoral allocation of water, water
conservation, improved water quality).
– Policies, legislation (e.g. water rights, water markets) and regulations may be modified to
benefit the poor directly or indirectly by providing livelihood assets.
– Community management of water and other natural resources may ensure that benefits
from the use of water resources are equitably distributed and the poor benefit from better
management of resources.
– Community management of water delivery services in rural and urban areas will ensure
the participation of communities in planning and management of delivery systems based
on demand responsive approaches (DRA) and ensure their financial and institutional
sustainability.
– Targeted subsidies and programs of service provisioning directly help the poor in securing
needed water supplies at prices they can afford. More efficient service delivery would
free-up financial resources that could be deployed for meeting the needs of partially served
and unserved poor.

4 INVESTMENTS IN WATER INFRASTRUCTURE AND POVERTY ALLEVIATION

Investment in water projects has provided livelihood, employment and incomes to the poor
people in developing countries. There are considerable multiplier benefits from irrigation and
power projects to farmers and agricultural workers in terms of increased employment and real
wages and income in agro-industries. Although this may not be true in some countries (e.g.
China, Spain) major water development projects in Brazil, India, Malaysia and the USA show
large direct benefits (from irrigation and hydropower) but with indirect benefits typically twice
as large (World Bank, 2003). In many cases the poor benefit enormously from this economic
204 R. Bhatia & M. Bhatia

activity. In Petrolina, in Northeast Brazil, for example, water infrastructure has been the basis
for the development of a dynamic rural economy. This has meant the creation of a large
number of high-quality, permanent agricultural jobs (40% held by women). And for every job
in agriculture, two jobs have been created in the supporting commercial and industrial sectors.
These opportunities have meant a reversal in the historic pattern of out-migration, with the
benefiting districts growing at twice the state average. Similarly in India, water infrastructure
has evened out the seasonal demand for labor, resulting in major gains for the poor. Irrigation
projects benefited, both directly and indirectly, by raising the demand for labor, a large number
of poor agricultural laborers who had no assets or training of any kind. For example, during its
operation over 30 years, the Bhakra dam in Northern India has transformed the lives of millions
of people in the states of Punjab, Haryana and Delhi, and provided seasonal employment to
hundreds of thousands of poor migrant workers from Bihar and Eastern Uttar Pradesh.

4.1 Poverty reduction impacts of Bhakra dam, India


The results of a multi-country World Bank study (for details: Bhatia et al., 2004) show that
investments in irrigation and hydropower projects have very significant indirect economic
impacts that could be as high as 90% to 100% of direct economic impacts.
The multi-sectoral, economy-wide models used for multiplier analysis also provide quan-
titative estimates of income distribution and poverty impacts of dams. The Social Accounting
Matrix (SAM)-based, fixed price, multiplier models provide income and consumption esti-
mates under With Project and Without Project situation for various household groups (rural
self employed, agricultural labor, marginal farmers, household quintiles in rural and urban
areas). These data have been used to compute gains or losses associated with the dam for
each category of household. The benefits to the poorest people (e.g. agricultural labor) from
irrigation and hydropower projects are sometimes higher than the benefits to other households.
Thus, investments in large water projects help significantly in the reduction of poverty in the
regions and beyond. For example, in the case of the Bhakra dam, it has been estimated that
agricultural labor gained a 65% increase in income as compared to a rural average increase of
38% under the With Project scenario compared with a hypothetical situation where the project
had not been undertaken (Figure 1)2 .
One of the major benefits of irrigation projects for the urban and rural poor is a substantial
reduction in prices of foodgrains3 . Given the high crop yields and substantial marketed surplus
of foodgrains resulting from surface and groundwater pumping (helped by recharge from
irrigation canals and power from the Bhakra dam), only two states (Punjab and Haryana)
provide around 70% of the all-India requirements of foodgrains distributed by the government
in fair-price-shops to the urban poor all over India (at relatively low food prices). The urban
poor have also benefited from surplus foodgrains from Bhakra distributed all over the country
through fair-price shops. Bhakra dam contributed 30 million tons or 60% of total foodgrains
procured by public distribution agencies in 2000–01. Remittances (around Rs 3548 million

2
Does not include the income of the 1 million workers who migrate seasonally from Bihar and Eastern
Uttar Pradesh each year.
3
This is evident from food prices in areas where crop yields are high (Bhatia, 2003). Higher marketed
surplus in irrigated areas provides additional foodgrains to the public distribution system that provides
foodgrains to the urban poor at prices they can afford.
Water and poverty alleviation: the role of investments and policy interventions 205

80
70
Percentage of increase
60
50

40

30
20

10
0
Landowners Agr. labor Rural non-agr. Rural others Urban
Types of households

Figure 1. Percentage change in income of different types of households With Bhakra dam.

Without Project With Project


3.500

3.000

2.500
Income (103 Rs)

2.000

1.500

1.000

500

0
Marginal Small Medium Large Workers
Households

Figure 2. Income of households With and Without Project 4 .

or US$ 75 million in 1995–96) sent by migrant labor working in the Bhakra command have
benefited millions of poor in the villages in Bihar and Uttar Pradesh, with resulting multiplier
and downstream effects.

4.2 Income distribution impacts of small check dams in Bunga village, India
In a recent case study (Malik & Bhatia, 2004), an attempt has been made to analyze the direct,
indirect and multiplier effects of the check dams constructed in Bunga village in Haryana.
Bunga has been a successful example of how community management of water and natural

4
Rs: Indian rupees. Rs 1 = US$ 0.02161 (September 1, 2004).
206 R. Bhatia & M. Bhatia

resources can benefit people, particularly the poor. In the Bunga village, a small village of 178
families (1100 persons), near Chandigarh in Haryana, India, the two check dams are providing
irrigation to about 276 ha of land.
The results show that the gain for land holding households (except for large farmers) from
the dam was relatively higher than the average difference of 48% between village-level incomes
under with and without project situations. In the case of Bunga check dam, the household
categories are landless workers, marginal farmers, small farmers and large farmers. The
increase in the incomes of marginal farmers was 50% as compared with the average increase
of 48% under With and Without Project situations. The increase in the income of small farmers
was 59% as compared to the average increase of 48% underWith andWithout Project situations.
Further, the worker households did not benefit much from the dam because the demand for
increased labor from irrigated crops was met from family labor by most of the farm families.
Due to relatively small farm holdings, the farm households did not increase their demand for
hired labor that would have benefited landless households. However, the worker households
benefited indirectly from the dam in terms of higher incomes from milk production and shops.

4.3 Poverty and irrigation


Based on an empirical study (Bhatia, 2003), it was found that irrigation was positively asso-
ciated with reductions in poverty levels in different states of India. The proportion of persons
above the poverty line is higher in states/regions where a higher percentage of cropped area
receives irrigation (e.g. Punjab, Haryana and West Uttar Pradesh). The empirical results based
on the analysis of data for 17 states for selected years and 81 agro-climatic regions (for 1984–
85) show that: (1) irrigation is positively associated with the level of per capita consumption
expenditure; (2) variations in real wages alone account for 60% to 70% of the inter-state vari-
ation in the proportion of population above the poverty line. Real wages, in turn, are influenced
by the level of irrigation in a state; (3) crop yields of rice and wheat are significantly higher in
states with higher proportion of area irrigated; and (4) wheat prices are significantly lower in
the states where the proportion of gross irrigated area to gross cropped area is relatively high.
Multiple uses of irrigation canals and tanks, in many countries, have provided opportunities
for the poor in terms of additional employment in crop production and income from livestock
and fishing. For example, in the Kirindi Oya Irrigation and Settlement Project (KOISP) in
Sri Lanka, significant benefits have been identified from crop production, fishing, tourism,
livestock grazing, fuelwood collection and shell mining (Meinzen-Dick & Bakker, 2001).
Despite many direct and indirect benefits to a large number of people (including the poor
and disadvantaged), a number of large water projects have resulted in adverse social and
environmental consequences. Some of these relate to sufferings of people (particularly the
tribals, native populations) due to absence of, or delays in, relief and rehabilitation of project
affected people (WCD, 2000). Environmental consequences relate to salinity and water logging
in large tracts in Pakistan and Western India, reducing cultivable land and affecting drinking
water supplies for people and animals.

4.4 Irrigation and returns to investment in education


Pritchett (2001) has found that for India, irrigation infrastructure has a major impact on returns
to investments in education. Returns to investing in five years of primary schooling compared
to no schooling in Indian districts where agricultural conditions were conducive to adoption
Water and poverty alleviation: the role of investments and policy interventions 207

of Green Revolution technologies, was as high as 32%. However, in districts where conditions
were not conducive to such irrigated agriculture, estimated returns to such schooling were
negative. At the same time, it has been found that in unirrigated districts in India 69% of
people were poor, while in irrigated districts this share drops to 26%. [It should be noted,
however, that information about these respective poverty shares before irrigation investments
took place, were not known].

5 POVERTY IMPLICATIONS OF IMPROVED MANAGEMENT OF WATER


RESOURCES

Improved management of water resources can help the poor people, both directly and indirectly,
in a variety of situations. Here we describe a few such situations where water resources
management can help in poverty alleviation/reduction.

5.1 Flexible water allocation and management benefit the poor


In a recent study (World Bank, 2004) for the Indian state of Tamil Nadu, it has been shown that
flexible water allocation among sectors provides substantial benefits in terms of higher incomes
for almost all the income groups. For example, the increase in overall income at the state level is
about 20% compared with a scenario of fixed water allocation. Some of the details of the study
on income distribution impacts of improved water management practices are given below.
This analysis of water allocation in the state of Tamil Nadu for the year 2020 has been done
under two alternative scenarios:

S-1. Fixed allocation


Business as Usual Scenario (BAU): Fixed sectoral water allocation based on proportion of
water allocated to agriculture, industry etc. as estimated in the year 2000. Additional supplies
of sustainable groundwater will also be allocated in the same proportion as in 2000.

S-2. Flexible water allocation


Volume of water available in this scenario is the same as in S-1 but water is allocated to all
sectors based on the estimated economic value of water in each sector (as defined by the
Willingness to Pay – WTP – for water in each sector). It is assumed that control structures
for water distribution among various users will be in place and incentives to transfer water
from one user to another will be present. This may require water rights, including rights
to transfer/sell water to other users as well as institutional interventions such as the water
users associations (WUAs). This will also include needed interventions in terms of transfer
and delivery mechanisms including the needed secondary and tertiary system improvement
options such as lining and on-farm development (OFD) works for effective water control. A
number of simulations have been carried out for this scenario by taking into account additional
water available from the Godavari link, water conservation in agriculture and industry and
removing the subsidy on the price of electricity used for water pumping.
The analysis is carried out in two steps (for details: World Bank, 2004). Estimation of direct
and indirect impacts of alternative strategies of water allocation on the economy has been
assessed through a two stage inter-related modeling process. In the first stage an Optimization
Model has been formulated to assess the direct impact of the increase in water demand on intra
208 R. Bhatia & M. Bhatia

and inter-sectoral allocation of water, and the direct impact such reallocation of water has on
these water using sectors. The second stage uses a Social Accounting Matrix (SAM)-based,
fixed price multiplier model of the Tamil Nadu economy to quantify the direct and indirect
impacts on other sectors of the economy and on the poverty levels under alternative scenarios.
The SAM-based model takes (as input to the model) the estimated output levels of sectors
directly affected by alternative scenarios of water allocation (e.g. optimum levels of output of
various crops, industries from the Linear Programming Model).
5.1.1 Income distribution aspects of water allocation
The SAM-based multiplier model disaggregates households into nine categories. Based on
the survey done in 1997, these nine income groups represent the following occupational
classification:
Rural household categories: (R1) rural self employed in non-agriculture; (R2) rural agri-
culture labor; (R3) rural other labor; (R4) rural self employed in agriculture; (R5) rural other
households.
Urban household categories: (U1) urban self employed; (U2) urban regular wage salary;
(U3) urban casual labor; (U4) urban other households.
In rural areas, R2 rural agriculture labor constitutes 45% of total rural population of 34
million people. Urban casual labor or U3 accounts for 19% of urban population, while U2,
urban regular wage earners, constitute the largest group at 43% of the total urban population
of 27 million.
It may be noted that under flexible water allocation (S-2), all income groups receive higher
incomes (except R2), when comparing with incomes received under fixed water allocation
scenario (S-1). On average, rural population gains 14% while R3 (rural other non-agriculture
labor households) gain 21% and rural non-agriculture households (R1) gain 19%. This is
because all these households gain indirectly from higher incomes generated in services, con-
struction, transport and communications and ancillary manufacturing. As a result of reduction
in direct incomes from farming, rural self-employed (farmers) gain only 3% while agricul-
tural labor households have incomes under solution S-2 that are marginally (2%) lower than
incomes under solution S-1.
It may be useful to compare per capita incomes across income groups for the years 2000 and
2020 (Figure 3) under Fixed and Flexible allocation. Over the two decades (2020 over 2000),
under the flexible water allocation strategy, overall per capita income increases by 500%. The
incomes of rural population increase by 280%, while those of urban people increase by 580%.
The incomes of the lowest income group in rural areas also increase (by 120%), although
much lower than that for other rural or urban groups. The lowest income group in urban areas,
however, gains 625% in per capita income, even higher than the average for urban population.
The income groups in the Figure 3, represent the following occupational classification as in
the survey done in 1997: (R1) rural self employed in non-agriculture activities; (R2) rural
agriculture labor; (R3) rural other labor; (R4) rural self employed in agriculture; (R5) rural
other households; (U1) urban self employed; (U2) urban regular wage salary; (U3) urban
casual labor; (U4) urban other households.

5.2 Community-based management of resources and poverty alleviation in India


Two villages in India (Sukhomajri and Ralegan Siddhi) have shown that it is possible to reduce
poverty by bringing about community-managed ecological regeneration (for details: Agarwal,
Water and poverty alleviation: the role of investments and policy interventions 209

Fixed Scenario Flexible Scenario


250
Per capita income (103 Rs)

200

150

100

50

0
R1 R2 R3 R4 R5 U1 U2 U3 U4
Classification of income groups

Figure 3. Per capita income of different groups under Fixed allocation and Flexible allocation
(year 2020).

1999). In both villages, development of transparent and participatory community-based


decision-making institutions, and establishment of community property rights over the local
natural resource base was critical. These institutions decided natural resource management pri-
orities, resolved conflicts within the communities, and determined burden- and benefit-sharing
rules. Technologically, the starting point was rainwater harvesting – centuries-old tradition of
Indian villages – which slowly led to the regeneration of the entire village ecosystem and the
associated rural economy.
The village of Sukhomajri, near the city of Chandigarh, has been widely hailed in India for
its pioneering efforts in microwatershed development. In 1979, when the nation was facing
a severe drought, the villagers built a small tank to capture the monsoon runoff and agreed
to protect their watershed, in order to ensure that their tank did not get silted up. Since then
the villagers have built a few more tanks and have protected the heavily degraded forest that
lies within and around the catchment of its minor irrigation tanks. The tanks have helped
to increase crop production by nearly three times, and the protection of the forest area has
greatly increased grass and tree fodder availability. This, in turn, has increased milk production.
With growing prosperity, Sukhomajri’s economy has undergone a change. The villagers have
replaced their thatch-and-mud dwelling with brick-and-cement houses, and most of the houses
boast of radio sets, electric fans, sewing machines and television sets. A survey conducted
in 1998 revealed that the income distribution in Sukhomajri, situated in an agriculturally-
depressed region, today matches the income distribution of rural Haryana which is one of the
most agriculturally prosperous states of India5 .
In 1975, Ralegan Siddhi was a village stricken by poverty (Hazare, 1997). It had hardly one
acre of irrigated land per family (1 acre = 0.4047 ha). Yield was less than 0.75 tons/ha. Food
production was only 30% of the village requirements, some 15%–20% of the families were

5
One of the most impressive savings resulting from the project is in the reduced cost of desilting the
Sukhna lake which supplies water to the downstream city of Chandigarh. The inflow of sediment has
come down by over 90%. This saves the government Rs 7.65 million (US$ 165,000) each year in dredging
and other costs.
210 R. Bhatia & M. Bhatia

undernourished and most men migrated each year to look for work. In 1979, a social worker
began work in the village by constructing storage ponds, reservoirs and gully plugs. Due to
the steady percolation of water, the groundwater table began to rise. Because of the increased
availability of irrigation water, land that was lying fallow came under cultivation and the total
area under farming increased from 630 ha to 950 ha. The average yields of millets, sorghum and
onion increased substantially. Every effort was made in the village to ensure equitable access
to the resources generated. This was particularly important when the conserved water was
still small in compared to potential demand. Water is distributed equitably. As cultivation of
sugarcane – common in the command areas of state-managed irrigation systems in the state –
requires a large quantity of water, it was forbidden in the early years in Ralegan Siddhi to
ensure that the limited amount of water available was distributed equitably to all farming
households. Only low water-consuming crops were allowed. All families got water in turn. No
farmer would get a second turn of irrigation until all families had been served. Even where
individuals had private dug wells, they were persuaded to share water with others. Water
conservation efforts resulted in increased availability of groundwater that in turn facilitated
the development of community wells. Water from these wells, supplied at a moderate price,
has enabled farmers to grow two to three crops a year including fruits and crops, some of
which are exported. A 1998 survey revealed that the monthly income distribution in Ralegan
Siddhi is much higher than the income distribution estimates of rural Maharashtra.

5.3 Community management of watershed and natural resources


There are interesting cases of cooperative watershed management (World Bank, 2003), stimu-
lated by the recognition of dam owners that upper catchment management is imperative for
maintaining the value of their assets. Thus, for example, the Nam Theun 2 hydropower project
in Laos provides support for communities to improve the management of the catchment. And
the private companies which operate the water concessions in Manila are similarly investing
heavily in catchment management to preserve the quality and quantity of the water on which
the city depends.
Another interesting case relates to the efforts to manage the watershed of the river La
Quebrada Chocho, Colombia, with the creation of an Association of Users to buy the land
around the watershed, reforest it, and preserve it in an integrated manner. Participatory diag-
nostic raised awareness of problems and built consensus on solutions such as tree planting
and the need for a solid management team supported by Participatory Action Research (PAR)
trainers. Training was useful to bring a common vision on watershed management, to solve con-
flicts, to build a management team. Decentralized community management brings ownership
of integrated water management.
In another case involving native populations in Mexico relates to civil society initiatives
to restore the traditional aztecan chinampas at the Xochimilco watershed. This case shows
the implementation of a common agreed project among different water users for a complete
hydrological regeneration of the chinampa area. It also gives examples of a large-scale experi-
ence of conflict resolution on watershed management, in an urban environment to devise a
plan of action, providing us with their experience of how they have blended traditional and
modern technology into a civil society’s initiative.
The NGO-led Shivalik hills watershed management project in India, financed by the
World Bank (2003), seeks to scale up the lessons from many successful NGO-led watershed
Water and poverty alleviation: the role of investments and policy interventions 211

management projects. The project aims at simultaneously reducing erosion, increasing ground-
water recharge, and improving the livelihood of the poor. The major investments are in
terracing, establishing small check structures in eroded ravines, planting vegetative cover
on denuded hills, building small dams, and digging wells and canals which can make better
use of the preserved water resources.

6 WATER AND SANITATION SERVICES FOR THE POOR

During the last two decades or so, access to safe drinking water and sanitation facilities has
been provided to millions of poor people through government programs and international
efforts of multilateral and bilateral agencies. The extent of water and sanitation services
actually available to the poor has yet to be evaluated (as discussed elsewhere in this chapter).
Still, millions of poor people in urban and rural areas have yet to receive basic facilities of safe
drinking water and sanitation. For example, South Asia accounts for 500 million of the world’s
poorest people with an income below one-dollar per capita per day. Of the total population
of 1313 million (1998), 300 million people do not have access to safe water today, millions
more get much less water than considered necessary for basic needs and personal hygiene.
If adequate investments are not made by 2025, the additional population of more than 500
million people will not get enough safe water, particularly in slums in mega cities. In the case
of sanitation, in South Asia alone, 920 million do not have access to sanitation; if investments
are not made, 1450 million will be without sanitation, causing health problems and loss of
output. Hence, new investments in the water sector are critical to the survival, well-being and
poverty alleviation of 1000 million people in South Asia.
The estimated investments in the water sector in India (Bhatia & Malik, 1999) would be of
the order of US$ 420,700 million over the next 25 years (undiscounted, at 1997 prices). It may
be noted that the bulk of the investments (39%) will be required for providing urban water sup-
ply, urban sanitation and urban wastewater treatment. Treatment of industrial effluents would
require an investment of US$ 68,800 million over the next 25 years. A little over a quarter of the
capital cost will be required for food security investments including surface irrigation, ground-
water pumping, drainage, watershed development and flood control. Investments of the order
of US$ 76,000 million (18%) would be required for rural water supply and sanitation services.
Although considerable progress has been made in increasing coverage levels, the sustain-
ability of such approaches and technologies have been questioned (Hansen & Bhatia, 2004).
It has been argued that financial and institutional sustainability of government sponsored
systems cannot be ensured due to lack of cost recovery and accountability. These government-
sponsored systems are characterized by supply/technology fixes and do not take into account
what services people want and what they are willing to pay for them. Lack of any notions
of opportunity cost of water in alternative uses has resulted in tariffs that are too low for
domestic and industrial users. Non-targeted subsidies have resulted in the rich people getting
the benefits of water services and subsidies in urban areas. Even for the relatively better-off
sections of society, uncertain and inadequate water supplies have resulted in high cost coping
strategies to meet their demands for water.
In many cities in the developing world, millions of the poor people living in unauthorized
colonies do not have access to adequate water supplies and sanitation services and continue
to be partially served or unserved. In the absence of adequate provision by the government
212 R. Bhatia & M. Bhatia

agencies, people have to do with inadequate and contaminated water supplies. The poor
get intermittent and unreliable water supplies from public standposts, while the rich, who
have house connections, use water lavishly and pay subsidized rates, much below the cost of
supplying water (for details for Delhi: Zerah, 2000). As a consequence, many poor people have
to pay high prices (almost ten times the prices paid by the rich) for getting water to meet their
basic needs of drinking and cooking and personal hygiene. Millions of poor people in rural
areas do not get adequate water supply and have to walk miles to get water for meeting their
basic needs. Such an arrangement is not sustainable and significant changes will be required
in the approaches to provision of water services to meet the objectives of poverty reduction.
Poor men and women do not have access to resources to lift them out of poverty. They could
do better if the resources were made available and they could control them to upgrade their
environments (Esrey & Andersson, 1999). If local communities own their systems, they will be
more responsive to demand from community members and be able to share and recover costs.
It is also important that when responsibilities for managing water systems are given to local
communities, financial resources and technical assistance should also be made available to the
local leadership. For example, in Tamil Nadu, India, in the initial stages, although responsibil-
ities for water supply to rural communities was transferred to panchayats, they did not get the
budgeted money for maintenance of these systems, and engineering staff was not transferred
to work under the panchayats. In addition, micro-credit facilities should be made available
to community organizations, NGOs and others to provide loans and guarantees. Small-scale
private sector organizations and groups (such as vendors, truckers) can provide cost-effective
services to urban poor and need to be helped by changing regulations and financial assistance
policies of banks and micro-credit institutions (deLucia and Associates, 2000).

6.1 Community managed rural water supply schemes – Equity and sustainability
issues in Swajal water supply scheme in Uttar Pradesh, India ( for details:
deLucia and Associates, 2001)
In an attempt to improve rural water supply and sanitation services in the state, the Government
of Uttar Pradesh piloted an alternative service delivery model-Swajal (Own Water). Swajal
is part of the Uttar Pradesh Rural Water Supply and Environmental Sanitation Project. The
Project aims at delivering sustainable health and hygiene benefits to the rural population
through improved water supply and environmental sanitation services.
The Swajal project envisages a decentralized decision making and demand responsive
institutional framework. The model consists of three organizations: the Village Water and
Sanitation Committees (VWSC) at the community level, the Support Organizations (usually
NGOs) at the district level, and the Project Management Unit at the centre. Selection of a village
to be covered under the project is done with the consent of the village community represented
by the VWSC. The VWSC and the village panchayats (body of elected representatives) need
not necessarily be the same entity. In some villages the two institutions could be the same,
while in others they could be different.
After the launch of the scheme, much is not known about the water supplied through the
standposts in these villages – the only source from which poor non-member households can
draw water. What has been the impact in terms of number of standposts now available, the
quantity of water that is made available through these standposts, the number of users per
standpost, the time spent by households in collecting water from these standposts, and the
implicit tariff paid by these households vis-à-vis those having a household connection. Have
Water and poverty alleviation: the role of investments and policy interventions 213

the poor households relying on standposts for their water needs become worst off in terms
of availability of water, time spent in collecting water as also the implicit tariff paid for a
unit of water both in terms of what they were paying prior to the scheme, as well as now
in comparison to the households having a piped water connection? Based on a short field
visit to two villages (near Dehradun, in June 2000), it seems that in the pumping scheme
where fuel costs are high, financial viability has been achieved at the cost of equity, and the
poor are paying for inadequate and unreliable water supply, while the rich are enjoying the
benefits of adequate water supply through house connections provided under the community
management. There are significant trade-offs between objectives of meeting basic needs of
the poor and financial viability of small community-managed water delivery systems. For
example, in the case of Swajal community schemes in Uttar Pradesh, India, cost recovery
for a diesel-based pumping system in a village required that each poor household pays Rs 10
(US$ 0.22) per household for getting water from a public standpost (compared with Rs 50 –
US$ 1.08 – per month for a house connection). However, given the objective of maximizing
cash revenues, the Village Water Committee, over time, increased the private connections
and reduced public standposts from 10 to 5 for about 50 poor households in the village. As a
consequence, the poor households have to spend a lot more time in procuring water from the
standposts that is generally inadequate for their needs.

6.2 Community participation in urban sanitation


For obtaining the health benefits of reduced infection from faecal-oral diseases, particularly
for the children, it is necessary to provide adequate sanitation, that includes hygienic, well-
maintained, easily accessed toilets that are used by all family members, and safe and convenient
disposal of wastewater. In 2000, the total number of urban dwellers lacking adequate sanitation
was between 850 to 1130 million, much larger than the 400 million lacking improved provision.
This means that the funds required to meet Millennium Development Goals for urban sanitation
are likely to be more than double of what has been estimated so far.
During the last two decades, the participation of communities and NGOs has resulted
in the provision of toilet and bath facilities to about 4 million urban slum dwellers through
community toilet complexes in India. In addition, community participation has helped in the
construction of private toilet facilities for about 1.5 million urban people in several countries
(for details: Bhatia & Bhatia, 2004; Bhatia & Hansen, 2004; Hansen & Bhatia, 2004) including
India, Bangladesh, Pakistan, Angola, Burkina Faso, Ghana, Brazil and Bolivia. Involvement
of communities provides incentives for selection of low-cost technologies, ensures efficient
management of services and leverages a part of the required capital funds.

6.3 Constraints to scaling-up of islands of success


6.3.1 Capital costs are substantial for scaling-up
However, the efforts of community participation have to be scaled-up by a factor of 100 times
if adequate sanitation services are to be provided to about 500 million urban poor and slum
dwellers over the next decade or so. Taking a conservative figure of US$ 20 per capita for
providing sanitation services in urban areas, the estimated capital requirements are around
US$ 10,000 million for the poorest in urban areas in Asia, Africa and Latin America. Besides,
such a scaling-up would require actions at a number of levels in the international organizations
and central and local governments in developing countries.
214 R. Bhatia & M. Bhatia

6.3.2 Public funds for capital costs constrain scaling-up


Although NGOs play an important role in organizing communities, the institutional structure
of NGOs does not enable them to raise capital funds from financing institutions. Further,
NGOs are not interested in obtaining loans to construct facilities and take financial risks.
Invariably, the efforts of the NGOs (as in India) in providing community toilet facilities are
constrained by the total funds available to the municipalities or local bodies for providing
sanitation services. For investments in private toilets, subsidies (20% to 60% of capital costs
in Ouagadougou, Burkina Faso) and loans (in Kumasi, Ghana) with the exception of Orangi
Pilot Project in Karachi, Pakistan are typical instruments.
Governments, local bodies, international lending institutions and bilateral aid agencies
have to understand and accept that far greater funds are required for meeting the Millennium
Development Goals of providing adequate sanitation than the resources available in the past.
Governments and local bodies should consider additional taxes on water use (e.g. in Burkina
Faso) and/or property taxes that are specifically marked and allocated to provisioning of
sanitation services to the poor. Financing for sanitation will require special funding provisions
as well as providing sufficient funds through poverty reduction support credits.
Additional funds have to be raised through financing mechanisms for leveraging resources
through private sector participation and/or through greater community resources. Innovative
institutional arrangements will be required so that NGOs can set-up companies that can raise
capital funds and manage financial risks. The design of financing mechanisms for leveraging
resources would require linking with domestic credit markets; emphasis on risk management;
pre-investment funds for project preparation and financial resources for efficient management.

6.3.3 Financial sustainability requires targeted subsidies to cover O&M costs


Given the limited ability of the poor to pay for sanitation services, the greatest challenge for
community sanitation systems has been low tariffs, generating revenues that are too low for cost
recovery and financial sustainability of existing or new services (Bhatia, 2004). For a long-term
sustainability of investments in urban sanitation, the community schemes have to raise revenues
from users who are too poor to contribute enough to cover even the Operation and Maintenance
(O&M) costs. Other sources of revenues have to be found to meet O&M costs and a part of
the capital costs. For example, in India, NGOs assisting communities can raise revenues from
their toilet complexes by renting space or selling biogas and thus reduce their dependence on
government funds and subsidies. However, this will require changes in policies and regulations
for NGOs and funding organizations such as the municipalities or international donors/banks.

6.3.4 Private toilet programs do not cover the poorest


In some community projects, most of the households that have invested in on-site sanitation
systems seem to belong to the middle class or poor but relatively better-off households (e.g. the
participants in the Orangi Pilot Project in Karachi, Pakistan; in Ouagadougou, Burkina Faso;
in Luanda, Angola; and in many towns in Brazil). For extending these programs, the subsidies
need to be targeted specifically towards the poorest sections, who will not be able to afford
the systems, even the low cost options. Such subsidies are in the nature of capital cost sharing
through subsidized loans. Cross-subsidization or direct support from the local governments
may be helpful in some cases. From the experience of many projects (e.g. Condominial system
in Brazil), it is important to recognize that cost recovery and subsidy rules must be set in a
clear and transparent manner.
Water and poverty alleviation: the role of investments and policy interventions 215

7 POLICY INTERVENTIONS FOR POVERTY ALLEVIATION

7.1 Legislation, water rights and water markets


Water legislation provides the basis for government intervention and action, and establishes the
context and framework for action by non-governmental entities; hence it is an important ele-
ment within the enabling environment. Specific water laws have been enacted in a considerable
number of countries, but some still lack a water resources law per se. Although references to
water resources may be found in the national legislation, these are often dispersed in a multitude
of sectorally oriented laws and may be contradictory or inconsistent on some aspects of water
resource usage (Solanes & Gonzalez-Villarreal, 1999; GWP-TAC, 2000a, 2000b).
Water rights can affect the poor in a variety of ways including: (1) granting of special water
rights to the poor that are not linked to land-ownership; (2) protection of traditional and custom-
ary water rights that are available to the poor, tribal and native populations; and (3) protection
of water rights of native populations by ensuring that federal or state governments do not give
away those water rights to companies using the water resources for irrigation/hydropower/
households/industry.
In Sukhomajri village in Haryana, India, the villagers created a new institutional arrange-
ment in 1981 in the form of a water users’ association evolving a commonly acceptable method
of distributing water (Chopra et al., 1990). A village society was formed with all households
from the village as members. It was laid down that the landless or those whose lands did not get
water due to high terrain, etc., also had a right to water and could sell it to those whose demands
for water exceeded their share. Water rights were established for all members of the society.
In case water distribution pipelines did not reach the land of any member, water-selling rights
were also established. On an average, about two water discharges were made per household
during Rabi (winter) season. While water storage dams, pumps and other equipment were
maintained out of water charges ranging from Rs 6 to 10 per hour (US$ 0.13 to 0.22 per hour)
of water discharge, the surplus of the society was used in laying additional pipelines, etc.
While the society does not make any profit in the distribution of irrigation water, the creation
of equal water rights is its prime institutional contribution. The scope of the village water
users’ association, meant initially for the distribution of irrigation water, was extended to the
management of a large number of common property resources (e.g. contracting forest land
from the government, managing fishing in the irrigation tanks).
Changes in policies and legislation have a substantial impact on opportunities for the poor.
For example (World Bank, 2003), in 1992 Mexico passed a new water law, which introduced
radical changes in the way in which water is managed. Of greatest importance in these changes
were giving the users much greater say in the management of water, and introducing tradable
water rights. In some areas the effects have been dramatic, with substantial reductions in the
(unsustainable) pumping of aquifers, and with water moving from traditional low-value crops
to new high-value crops. Each drop of water now generates much higher economic returns
and each hectare of land and each drop of water now generate a direct demand for more than
twice as much agricultural labor (and therefore opportunities for the poor).

7.2 Targeted subsidies, water tariffs and pollution charges


According to the principle of managing water as an economic and social good, the recovery
of costs should be the goal for all water uses, unless compelling reasons indicate otherwise.
216 R. Bhatia & M. Bhatia

Yet, this principle entails inherent difficulties: how can principles of equitable access to water
used for basic human needs be taken into account at the same time? At a minimum, full supply
costs should be recovered in order to ensure sustainability of investment and the viability of
service providers. However, in many situations, even the achievement of this objective may
require direct subsidies for years to come.
Poverty alleviation policies might be incompatible with abrupt implementation of full
supply cost recovery in, for instance, some surface irrigation systems. In the provision of
municipal and rural water supply there are well-established practices of cross-subsidization
from better-off water users to the poor. The use of cross-subsidies does not necessarily
compromise the financial sustainability of utilities but they distort prices and patterns of
demand. For management purposes such subsidies should be made in a transparent man-
ner and, where possible, direct subsidies are the preferred option to reduce distortions in
the system. Under normal circumstances industries should meet at least the full economic
cost of the supplied water (Bhatia & Falkenmark, 1993; Bhatia et al., 1994; Bhatia et al.,
2000).
Chile has been able to implement a well-working system of focal subsidies in the water
and sanitation sector (Briscoe et al., 1997; GWP-TAC, 2000a). The success of the system
depends on the concerted effort and institutional capabilities of the national government, the
municipalities and the water companies. Other countries in Latin America have attempted
to replicate the very successful Chilean experience. However, the funds available did not
match the needs of the users, neither did the institutional capability of governments match the
monitoring requirements of system implementation and enforcement. For this reason some
countries, such as Argentina, have resorted to traditional cross-subsidies, despite the obvious
drawbacks of the system. The lesson is that before suggesting either focal or cross-subsidies,
countries and financing institutions should ensure not only financial and economic viability,
but also that institutional structures do allow efficacious implementation.

7.3 Special programs for financing connection costs


Maylinad Water Services Inc. in Metro Manila has a program called Bayan Tubig (Water for the
Community) under which the company provides individual connections to poor households.
Although water tariffs are the same for all households, the increasing block tariffs mean low
tariffs for households using 10 m3 per month or less. For poor households there are schemes for
payment of connection charges (US$ 60) in 12 monthly installments. In squatter settlements,
where space is a problem, water lines are run through neighborhoods and connected to a
battery of water with each meter assigned to a nearby home. From the meter, each homeowner
makes his own plastic connection, aboveground. Around 60,000 such new connections (50%
of total new connections) have been made under this scheme supplying water to more than
400,000 poor people.

7.4 Water tariffs


Water tariffs provide little incentives for the sustainable use of water if charged at a flat
rate independent of the amount used. In such cases, setting the right fee structure, imposing
progressively higher unit cost prices on high-volume users, may induce the more judicious
use of the resource, although the level of demand reduction will depend upon the nature
of the high-volume users. Such a structure also contributes to the financial sustainability
Water and poverty alleviation: the role of investments and policy interventions 217

of water authorities and to covering the cost of administering water resources management.
(GWP-TAC, 2000a).
It has been argued (Bhatia et al., 1994; Bhatia & Kijne, 1994; Rogers et al., 1998; Bhatia
et al., 1999) that sustainable and efficient use of water require the tariff to match not only
costs of supply (i.e. O&M and capital costs), but also opportunity costs, economic externality
costs, and environmental externality costs. Very often the tariffs do not even meet the full
supply costs, and sometimes the value of water is lower than the cost of supply! The evidence
is presented from the Subernarekha and Yamuna basins in India. For example, a comparison
of costs and values associated with urban and industrial water use in the Subernarekha river
basin is given in Rogers et al. (2002). The full cost of water to be US$ 0.47/m3 . The value
in industrial use was US$ 2.60/m3 ; the value in urban use was US$ 0.25/m3 . The industrial
tariff is more than double the urban tariff, but still incredibly low compared to the cost and
value of water. So the urbanites over-consume and value the resource less. In this case there is
an industrial to domestic to agriculture subsidy. In each case the agricultural tariffs are much
smaller than cost of supply in both basins.

8 CONCLUSIONS

To recapitulate:
– There is ample evidence to show that investments in water infrastructure (irrigation and
hydropower projects, small check dams) have considerable multiplier (direct and indirect)
benefits to poor farmers and agricultural labor in terms of increased output, employment
and real wages.
– Improved management of water resources (through flexible allocation among sectors
and users) provides benefits to the poor and can be an important instrument of poverty
alleviation/reduction in developing countries.
– A review of experiences in community management of rural water supply services points
to the need for careful monitoring of the impact of some community schemes on the poor,
particularly those where it is mandatory to recover, from the users, the high cost of pumping
schemes.
– Community participation in urban sanitation in the context of the Millennium Development
Goals should be encouraged. However, there are significant constraints in scaling up these
islands of success in terms of sources of funds, cost recovery and financial viability. Public
funds will be required to provide targeted subsidies to the millions of poor who can not
afford to pay for water and sanitation services. Leveraging of community funds and private
financial resources will require innovative financing mechanisms.
– To ensure protection and augmentation of water resources and their equitable distribution,
innovative practices will be required in legislation and regulation, water rights, pricing of
water and cross-subsidies. Financial sustainability of water investments will be necessary
so that the poor get the water and sanitation services they need and deserve.

REFERENCES

Agarwal, A. (1999). Population and Sustainable Development: Some Exploratory Relationships. Centre
for Science and Environment, New Delhi, India.
218 R. Bhatia & M. Bhatia

Agarwal, A. & Narain, S. (1980). Strategies for the involvement of the landless and women in afforest-
ation: five case studies from India (mimeo). World Employment Programme, International Labour
Office, Geneva, Switzerland.
Agarwal, B. (1995). Gender, Environment and Poverty interlinks in rural India: regional variations and
temporal shifts, 1971–1991. UNRISD Social Development News, 13.
Agarwal, B. (1997). Gender and Legal Rights in Landed Property in India. Institute of Economic Growth,
Delhi, India. Draft.
Bhatia, M. (2004). Financing and Financial Viability of Community Sanitation Projects: A Review of
Sulabh and SPARC Programs in India. Draft.
Bhatia, R. (2003). Poverty and Irrigation in India: Some Empirical Findings. Research Report, Resources
and Environment Group, New Delhi, India.
Bhatia, R. & Bhatia, M. (2004). Financing and Financial Viability of Community Sanitation Projects:
a review of Sulabh and SPARC Programs in India. Resources and Environment Group, India.
Draft.
Bhatia, R. & Falkenmark, M. (1993). Water Resource Policies and the Urban Poor: Innovative
Approaches and Policy Imperatives in Water and Sanitation. CURRENTS, World Bank. (Paper
presented at the Dublin Conference on Water and Environment).
Bhatia, R. & Hansen, S. (2004). Millennium Development Goals Challenges: What Can We Learn From
the Experience of Community Participation in Urban Sanitation Programs? Paper prepared for the
Norwegian Ministry of Environment.
Bhatia, R. & Kijne, J. (1994). Conflicts in Water Use: Sustainability of Irrigated Agriculture in Develop-
ing Countries. Proceedings of the Stockholm Water Symposium, August 9–12, 1994, Stockholm,
Sweden.
Bhatia, R. & Malik, R.P.S. (1999). Financing Needs of the Water Sector in India. Resources and
Environment Group, New Delhi, India. Draft.
Bhatia, R.; Rogers, P.; Briscoe, J. & Sinha, B. (1994). Water Conservation and Pollution Control in Indian
Industries: the Role of Regulation, Water Tariffs and Fiscal Policies in India, Water and Sanitation.
CURRENTS, World Bank.
Bhatia, R.; Rogers, P. & de Silva, R. (1999). Water is an economic good: how to use Prices to Promote
Equity, Efficiency, and Sustainability. Paper presented at the Stockholm Water Symposium.
Bhatia, R.; Kumar, R.; Misra, S. & Robins, N. (2000). Full Cost Pricing of Water – Options and Impacts:
A Case Study of the Impacts of Moving to Full Cost Pricing on Freshwater Demand, Recycling and
Conservation at the Tata Steel Company, India. UNIDO-IIED. Draft.
Bhatia, R.; Scatasta, M. & Cestti, R. (2004). Indirect Economic Impacts of Dams: Methodological Issues
and Summary Results of Case Studies in Brazil, India and Egypt. Vol. I–II, Forthcoming, World Bank,
Washington, D.C., USA.
Bos, E. & Bergkamp, G. (2001). Overcoming Water Scarcity and Quality Constraints. Water and the
Environment, IFPRI, Washington, D.C., USA.
Briscoe, J.; Salas, P.A. & Peña, T.H. (1997). Managing water as an economic resource: Reflections on
the Chilean experience. World Bank, Washington, D.C., USA. Draft.
Chopra, K.; Kadekodi, G. & Murthy, M.N. (1990). Participatory Development: People and Common
Property Resources. Sage Publications, New Delhi, India.
deLucia and Associates (2000, 2001). Best Practice Policy and Financing Approaches for Small-Scale
Infrastructure Service Providers in South Asia. Report prepared for the World Bank, October 2000
and March 2001.
Esrey, S. & Andersson, I. (1999). Poverty-Environment Interventions in Water and Sanitation: Key Issues
and Policies. UNDP-UNICEF.
GWP-TAC (2000a). Integrated Water Resources Management. Global Water Partnership, Technical
Advisory Committee. TAC Background Paper 4. 71 pp.
GWP-TAC (2000b). Letter to my Minister. By Ivan Chèret. Global Water Partnership, TechnicalAdvisory
Committee. TAC Background Paper 5. 30 pp.
Hansen, S. & Bhatia, R. (2004). Water and Poverty in a macro-economic context. Paper commissioned
by the Royal Norwegian Ministry of the Environment.
Hazare, A. (1997). Ralegan Siddhi: a veritable transformation. Ralegan Siddhi Pariwar Publications,
Ralegan Siddhi, India.
Water and poverty alleviation: the role of investments and policy interventions 219

Malik, R.P.S. & Bhatia, M. (2004). Indirect Economic Impacts of Check Dams in the Bunga Village,
India. In: R. Bhatia, M. Scatasta & R. Cestti, Indirect Economic Impacts of Dams: Methodological
Issues and Summary Results of Case Studies in Brazil, India and Egypt. Vol. II, Chapter 3. Forthcoming,
World Bank, Washington, D.C., USA.
Meinzen-Dick, R. & Bakker, M. (2001). Irrigation systems as multiple-use commons: water use in
Kirindi Oya, Sri Lanka. Irrigation and Drainage Systems, 15(2): 129–148. Also published as EPTD
Discussion, Paper 59. International Food Policy Research Institute (IFPRI), Washington, D.C., USA.
Meinzen-Dick, R.; Swallow, B.M.; Jackson, L.A.; Williams, T.O. & White, T.A. (1997a). Multiple
Functions of Common Property Regimes. IFPRI Environment and Production Technology Workshop.
Summary Paper 5. Washington, D.C., USA.
Meinzen-Dick, R.; Brown, L.; Feldstein, H. & Quisumbing, A. (1997b). Gender, Property Rights, and
Natural Resources. World Development, 25(8): 1303–1315.
Pritchett, L. (2001). Where has all the Education gone? World Bank Economic Review, 15: 367–391.
Oxford University Press.
Rees, J. (2001). The Risk and Integrated Water Resources Management. Paper prepared for the 9th Inter-
national Conference on the Conservation and Management of Lakes. Lake Biwa, Japan, November,
8–16.
Rogers, P.; Bhatia, R. & Huber, A. (1998). Water as a social and economic good: how to put the principle
into practice. Global Water Partnership, Technical Advisory Committee, Paper 2.
Rogers, P.; Bhatia, R. & de Silva, R. (2002). Water is an economic good: how to use prices to promote
equity, efficiency, and sustainability. Water Policy, 4: 1–17.
Solanes, M. & González-Villarreal, F. (1999). The Dublin Principles for water as reflected in a compara-
tive assessment of institutional and legal arrangements for integrated water resources management.
Global Water Partnership, Technical Advisory Committee, Paper 3.
Soussan, J. (2001). Poverty & Water Security. Paper prepared for the Water and Poverty Initiative, Global
Water Partnership, Stockholm, Sweden, December 2001. Draft.
Sullivan, C. & Meigh, J. (2003). Considering the Water Poverty Index in the context of poverty alleviation.
Water Policy, 5(5): 513–528.
Sullivan, C.A.; Meigh, J.R. & Giacomello, A.M. (2003). The Water Poverty Index: Development and
application at the community scale. Natural Resources Forum, 27(3): 189–199.
UN (United Nations) (1999). The State of world population. United Nations Population Fund.
UN (United Nations) (2004). Interim Report of the Task Force 7. United Nations Millennium Project.
WCD (World Commission on Dams) (2000). Dams and Development: a new framework for decision-
making. The Report of the World Commission on Dams. Earthscan Publications, UK.
World Bank (2003). Water Resources Sector Strategy – Strategic Directions for World Bank engagement.
Washington, D.C., USA. Draft.
World Bank (2004). Tamil Nadu Water, Economy and Poverty Study. A Draft Report by the Study Team,
World Bank, New Delhi, India.
Zerah, M. (2000). Water – Unreliable Supply in Delhi. French Research Institutes in India.
CHAPTER 13

Do investments and policy interventions reach the poorest


of the poor?

C.A. Sullivan
Centre for Ecology and Hydrology, Wallingford, UK

ABSTRACT: Macroeconomic policy instruments are often held to be the most effective tool for
poverty alleviation. In the water sector, various approaches have been used. These include flexible
water allocation schemes, water trading and transferable rights. Illustrations of the benefits of cross-
subsidization schemes are cited, and the need for private sector involvement in water distribution is
frequently stressed. In contrast to this, the case is made in this chapter that in addressing the needs of
the poorest of the poor, such measures are far from ideal. Weaknesses are highlighted both in terms of
theoretical frameworks and in terms of procedural failure, and it is due to these, it is argued, that poverty
continues to be ubiquitous throughout the world. How development is measured, and how benefits from
water infrastructure are assessed, are brought into question, presenting an alternative explanation of why
poverty is such a difficult ill to cure.

Keywords: Poverty alleviation, Water Poverty Index, investment, water policy

1 INTRODUCTION

The chapter Water and poverty alleviation: the role of investments and policy interventions
(Bhatia & Bhatia, this volume) provides an excellent review of water projects, highlighting
potential approaches that could be used as best practice tools for poverty alleviation. Many
examples are cited, from India and other parts of the world, demonstrating the practical
benefits from community participation, flexible water allocation schemes, property rights,
water trading, etc. The benefits of community water management are discussed on a number
of levels, with examples from many places to illustrate how it can work in practice. Finally, the
chapter concludes that while clearly investment in infrastructure will have a positive and more
than proportionate benefit for poor people, there is need for more involvement by the private
sector to finance it, as one of the problems of reaching the very poor is the fact that they simply
do not have the cash to support any kind of water provision system in a sustainable way.
The flexible allocation models described in the chapter do demonstrate how cross-
subsidization can take place at all scales. If one sector bears some extra costs as a result
of reallocations, it may mean that other sectors gain, and supplies are maintained to the poor,
or for domestic use. The additional cost to some sectors is compensated for, in pareto terms,
by the marginal benefits of extra water access accruing to the poor, and other, possibly more
productive sectors. This shows how the distributional impacts of cross-subsidization can be
spread widely throughout the economy, along with the potential to use water in a more cost
effective way, assuming allocations are made on the basis of efficient returns to water use.
222 C.A. Sullivan

Of course, it must be said that the determination of what is most efficient in any location will
be determined by social, cultural and political values, so representative consultation would be
essential if equitable decisions and financial instruments were to be put in place.
The study from Tamil Nadu state in India does suggest however that under a flexible alloca-
tion strategy, although overall incomes would increase significantly, more benefits accrue to
urban communities than to the rural population, who make up some 34 million people in that
state. Of these, 45% are classified as rural labour, and according to the analysis presented, these
are the group who benefit least of all from this approach, actually losing 2% of their household
incomes under a flexible allocation scheme. In reality therefore, this means that while overall
per capita income is estimated to increase by 500% between 2000 and 2020, for those who do
not manage to move to other forms of employment, the rural labouring poor experience a less
than proportionate rise in incomes, effectively subsidizing the benefits gained by others.
While this study does provide an interesting insight into both the methodologies of liveli-
hood analysis, and of how water management may impact on different groups, it does also
highlight some of the weaknesses of certain widely used techniques. In the analysis presented,
“water is allocated to sectors on the basis of the estimated economic value of water in each
sector, as defined by Willingness to Pay (WTP) for water in each sector”. There is much
literature on the issue of valuation of environmental goods and services (Farber & Costanza,
1987; El Serafy, 1991), and the work by Costanza et al. (1997) on ecosystems highlighted the
uncertainty associated with such values. The WTP methodology itself is fraught with pitfalls,
and to some extent at least, is an unreliable numeraire for policy formulation (Friend, 1993;
Hueting et al., 1998). The whole debate of how water is valued (Briscoe, 1996; Faucheux &
O’Connor, 1998; Spash, 1998), what it is used for, and who are the beneficiaries, are recurring
issues in water management, and the rest of this chapter is used to raise questions about how
we judge development impacts, and how the benefits of water allocation are actually assessed.

2 ADDRESSING WATER NEEDS

Water is a fundamental component of every ecological system. The development of policies


and tools which recognize the role of water as part of our life support system is essential for
the achievement of global security. In Africa, Asia and Latin America, some 70% of human
populations currently live in fragile ecosystems, where significant disruption to ecological
services could bring about what effectively would be irreversible consequences (Raskin et al.,
1997). This can already be seen globally in terms of increased rates of desertification, soil
salinization and deforestation, and in many areas of the world, such as the Sahel, Amazonia,
the Aral Sea, etc., where the impact of human activities has already begun to disrupt the life
support functions of those systems (UNESCO, 2003).
At the human scale, this situation is even more pressing, as currently, at least 1100 mil-
lion people lack access to an improved water source, and 2400 million lack access to basic
sanitation. As a result, every day some 6000 children in developing countries die for want of
a clean water supply and sanitation (WHO-UNICEF, 2000), and thousands more suffer from
the impacts of ill-health. In general, water scarcity negatively impacts on food security and
livelihood choices, and consequently, economic progress is stifled. Furthermore, educational
opportunities for poor families are often missed, and women continue to bear an unjust burden
of water provision, particularly in Africa and Asia.
Do investments and policy interventions reach the poorest of the poor? 223

Estimated daily water requirements for food


production, Km3
1,400

Daily water requirements,


1,200

cubic kilometres
1,000 Sub-
800 Saharan
600 Africa

400 South Asia


200 Latin
America
0
2000

2010

2020

2030

2040

2050
Figure 1. Increasing demand for water as human populations rise.

The impact of human population growth is a major issue when considering the future
challenges for water management (Falkenmark, 1990). If we examine the effect that this will
have on demand for water for both domestic use and food production, we can see that expected
rises in demand for these uses are significant (Postel, 1998). This is illustrated by Figure 1,
which shows the likely increase in water for food production over the next 50 years. This
is based on a water requirement for food production of 1400 L/d of water per person, with
population growth rates maintained at the 1999 level. It is important to note however that these
figures make no attempt to incorporate any increases which will also occur as a result of rising
standards of living.
Since the overall amount of water available for human use is limited by time and space,
it is reasonable to assume that as demand increases following population growth, per capita
resource availability shrinks, generating conflicts over use. Gleick (2000) has examined many
aspects of water resources and entitlements, especially with respect to global security, and
today, the issues of both poverty and water are now attracting considerable attention from a
security point of view. The widespread publication of global disparities in water accessibility
in such meetings as the World Summit on Sustainable Development in Johannesburg in 2002
and the 3rd World Water Forum in Kyoto in 2003 have emphasized the need to address the
problem of water management more effectively, both at a local and international scale.

3 RECONSIDERING THE LINKS BETWEEN DEVELOPMENT AND


ECONOMIC GROWTH

In the past, economic development has in most cases taken precedence over all other issues
(Lumby, 1979), but more recently, attention has been drawn to the need to expand the way
in which we view the development process (Lipton, 1988; Sen, 1995). Numerous examples
can be found where ecological disruption has resulted from development projects designed to
increase agricultural or industrial production. These have occurred because knowledge of the
complexities of ecosystems is limited, and values of the relevant environmental attributes have
been ignored (Jacobs, 1997). Compounded by a scientific approach which has been focused
rather than holistic, to some extent at least, this has led to unrepresentative theories of growth
224 C.A. Sullivan

economics. These theories, on which many development projects are founded, are based on
understandings which:

(a) suggest that man-made and natural capital can infinitely be substituted, and
(b) ignore the constraints on production from the laws of thermodynamics (Daly, 1999).

Clearly, while it can be shown that some measure of economic capital is generated from the
depletion of natural resources, it can also be easily shown that certain natural resources cannot
be reproduced by utilization of financial or physical capital (Kaufmann, 1995). This refutes
the concept of perfect substitutability of factors of production, which is a basic assump-
tion underlying the positions held by eminent economists such as Julian Simon (Simon &
Khan, 1984) and Wildfred Beckerman (1995). Furthermore, the fact that money generated by
exploitation of natural capital is accounted for in terms of income streams rather than capital
depletion, brings about an inevitable undervaluation of such resources, and consequent policy
failure.
The existence of entropy, as explained by the laws of thermodynamics, means that even
the most efficient production system must produce waste, thus making it clear that the idea
of infinite resource recycling and substitution is physically impossible. The failure of early
growth theories to take account of these real world conditions is one of the reasons why many
water projects developed in the past have failed to live up to expectations, and why numerous
examples exist of inequitable, and perhaps unexpected, development outcomes.

4 THE RELATIONSHIP BETWEEN WATER USE AND ECONOMIC


DEVELOPMENT

At a national level, it can be seen that countries which have higher levels of income tend to
have a higher level of water use, as can be shown in the examples in Table 1.
If however we look more closely at the returns from water use by sector, we can see that
in some countries, different sectors play different roles in economic performance, and the use
of water in these sectors may be a determining factor in the progress of an economy. Table 2
shows sectoral correlation coefficients for a selection of rich and poor countries currently
facing some degree of water shortage. In this example, these countries have been identified

Table 1. Water use and national income.


Annual water withdrawals per capita, m3 (1970–1987)

GDP per capita, Industrial and


US$ (1990) Domestic agricultural Total

Tanzania 110 8 28 36
Sri Lanka 470 10 493 503
South Africa 2530 65 339 404
United Kingdom 16,100 101 406 507
Sweden 23,660 172 307 479
United States 21,790 259 1903 2162

Source: World Bank (1992: Development and the Environment, Tables 1, 33).
Do investments and policy interventions reach the poorest of the poor? 225

as having a per capita Gross Domestic Product (GDP) of either less than, or more than US$
10,000 (on purchasing power parity rates), and the degree of water shortage is indicated by
having less than 0.5 km3 /yr per capita freshwater withdrawals.
From this information, we can see that in the agriculture sector in poorer countries, there is
some correlation between agricultural water use and the contribution of agricultural outputs
to GDP. For industry, in poor countries, higher levels of industrial water use tend to generate
greater contributions to GDP, whereas there is no such correlation in rich countries. While
these correlations provide no suggestion of a causal relationship, they do suggest that in poor
countries, when water is used for industry, the economic returns are of greater importance to
the nation than in richer countries. It also reflects the fact that in higher income countries,
the service sector (which is not a heavy water user) is much more important than it is in poor
countries. These figures also indicate that in poor countries, there is a high negative correlation
between the contribution of agricultural activity to the economy, and the country’s score on
the Human Development Index (HDI), reflecting the fact that a high level of agricultural
dependence may be associated with lower levels of economic and social wellbeing. This
confirms the fact that developing economies heavily dependent on agriculture tend to be price
takers on world commodity markets, often getting very little return for the use of any of their
resources, including water.
From this sample of countries, there also appears to be a stronger link in the higher income
group between per capita water use and per capita GDP, than in the lower income group, again
suggesting that the richer countries are in a better position to get higher benefits from their
water use than do poor ones. While this may reflect the fact that they have better control of
other factors of production, this is clearly an area of research which should be explored much
further, and while these figures can only be taken as illustrative, they do suggest that how
water is managed within an economy can have a direct impact on the economic welfare of
society. It is important to note, however, that these figures take no account of rainfed farming
systems that provide significant amounts of food throughout the world, much of which may
be unaccounted for in national accounts due to home consumption. Furthermore, no account
is taken in these figures of environmental damage associated with any form of water use or
allocation.

Table 2. The relationship between economic performance and water use.


to value added to GDP

to value added to GDP


agricultural water use
(%) and contribution

(%) and contribution


Correlation between

Correlation between

Correlation between

Correlation between

Correlation between
value added to GDP
industrial water use

water withdrawals

water withdrawals
from agriculture

from agriculture
and HDI score

and HDI score


from industry

and GDP per


capita value

Countries with high −0.03 −0.07 −0.06 0.51 0.47


GDP per capita
Countries with low 0.21 0.69 −0.51 0.11 0.27
GDP per capita

Note: Both groups of countries suffer from some degree of water shortage.
226 C.A. Sullivan

5 THE NATURE AND SCALE OF POVERTY

Poverty is a multidimensional concept, and numerous definitions of it have been provided in the
vast literature that has been generated on the subject. Recurring themes in the literature include:
– Poverty in the context of development (van der Gaag, 1988; Sen, 1995; UNDP, 2000);
– Poverty measurement (Lipton, 1988; Desai, 1995; World Bank, 1996);
– Poverty thresholds (Orshansky, 1969);
– Poverty and gender (Rosenhouse, 1989);
– Poverty and welfare (World Bank, 1998);
– Poverty and food (Malseed, 1990);
– Poverty and politics (Uvin, 1994);
– Poverty and health (WHO, 1992);
– Poverty and vulnerability (Sullivan et al., 2002; Meigh & Sullivan, 2004).
One consensus that certainly exists in all this literature is firstly that there are too many
poor people, and often they find themselves trapped in poverty by circumstances beyond their
control.

6 HOW ECONOMISTS MEASURE POVERTY

Methods currently in use to assess poverty need to be considered when discussions of poverty
alleviation take place. There are a number of approaches used to quantify poverty levels, includ-
ing the Poverty Line, the Headcount Index, and the Poverty Gap. All of these approaches are
based on national income figures, and as averages, are not very representative of regional
variations. As a result, they often fail to accurately represent the levels of poverty experienced
in different communities. Importantly, measures of per capita income are recognized to be
inadequate to represent human wellbeing. While money measures may provide some means of
comparison of economic activity, they take no account of non-monetary attributes of human
wellbeing, nor of the value of women’s household labor, nor indeed of depreciation of natural
capital.

7 ABSOLUTE POVERTY

In absolute terms, there is always a minimum level of certain resources that people need in
order to survive, and without these, a person will find himself in a state of poverty. Access
to water resources is likely to be a major determinant of absolute poverty. Absolute poverty
describes poverty “without reference to social context or norms and is usually defined in terms
of simple physical subsistence needs but not social needs”. Furthermore “absolute definitions
of poverty tend to be prescriptive definitions based on the assertions of experts about people’s
minimum needs” (Gordon & Spicker, 1999). The World Bank employs an absolute definition
of poverty in the form of a poverty line. The poverty line, or minimum income required, is
deemed as US$ 1 per day, and more than 1000 million people in the world today fall below
this level. While such a definition does serve to quantify the extent of poverty in a monetary
sense, it does not encompass important social factors, such as being free of obligations to make
Do investments and policy interventions reach the poorest of the poor? 227

100
RESOURCES
80

ENVIRONMENT 60
ACCESS
40

20

USE CAPACITY

Awarakotuwa Tharawaththa Agarauda Tissawa

Figure 2. Displaying findings from the Water Poverty Index (Sullivan & Meigh, 2003).

children work for others, or the ability to decline demeaning jobs. These are two examples of
the elements of poverty that are often identified by the poor themselves. As a result of these
omissions, many measures may fail to accurately indicate the levels of poverty experienced
in different communities. Importantly, measures of per capita income are now recognized to
be inadequate to accurately represent human wellbeing (UNCED, 1992; Scoones, 1998).

8 LINKING WATER AND POVERTY

In an attempt to address the inadequacy of many different water evaluation techniques, new
approaches have been developed, and one of these, the Water Poverty Index (Sullivan, 2002;
Sullivan et al., 2003), tries to capture the diversity of the many facets of water, and its links
to wellbeing. This is an indicator based approach, and Figure 2 illustrates how the results of
the Water Poverty Index (WPI) can be displayed, delivering complex information to policy
makers in a simplified way. Using this approach, it is not only possible to identify the spatial
variability of water provision, but also the variation in component values is shown, highlighting
strengths and weaknesses in particular locations. The use of this approach does facilitate the
inclusion of a greater variety of issues, and increases the accountability of water management
decisions. Using this approach, decision makers are able to take account of many aspects of
water management, to distinguish between locations, and prioritise investments.

9 ACCOUNTABILITY AND UNACCOUNTABILITY

Accounting for water is an important challenge underlying efficient and equitable manage-
ment. Unaccounted-for water is a term often heard in discussions with water professionals,
representing as it usually does, an apparently unrecoverable cost. This has long been an area
228 C.A. Sullivan

of attention for executives of water companies, and management is encouraged to reduce its
level by whatever appropriate means. While this may be admirable from a business efficiency
point of view, it is often not equitable, as in many instances, those currently benefiting from
unaccounted-for water are the poorest in society, and least able to pay.
In economic analysis, there are often many activities which are unaccounted-for. These
would include obvious things like crime and the proceeds from it, but also other things like
charity work, and unpaid housework. Often, another unaccounted-for aspect of economic anal-
ysis is the environment. Environmental costs associated with production are almost always
unaccounted-for, and there is much need to reassess the processes by which we evaluate devel-
opment options when these unaccounted-for dimensions create such uncertainty (Hannon,
1991; Jacobs, 1994; Daly, 1999). Furthermore, accounting systems are designed on the basis
of standards established in developed economies. Sometimes this creates problems, as activi-
ties and values particularly relevant in some places are not relevant in others (Opschoor, 1998).
The question of whose values should be counted is particularly important in considering water
provision.
This leads us to the point of unaccounted-for people. These could include the millions
of people who over the last five decades have been ousted from their homes in the interest
of policy implementation and water management (Dreze et al., 1997). It could include the
millions who live in the squatter settlements fringing almost all the world’s large cities. In the
case of these people, they are often unaccounted-for, since they lack any property rights or
tenured access. Unaccounted-for people would also include those millions who are simply not
registered in any formal way in the population records. Even in the most advanced places, data
is often missing. So in megacities across the world, many people are unaccounted-for, simply
because that is what they want to be. This failure to account accurately for human numbers is
another reason why many attempts to reduce poverty are often flawed.
Another barrier to the achievement of better water access is due to the widespread imple-
mentation of water pricing. While there is much logic in this development, there is little doubt
that large numbers of the world’s poor are simply not in any position to pay anything for water
at all. At the same time, it is known that in many countries, the poor pay more for water than
the rich, (as shown in Table 3), and this is often given as justification of why more streamlined
pricing should be adopted.
While some examples of small scale cross-subsidization are to be found in many places,
there is little doubt that much more should be done to implement more such schemes to address
this. It could be argued that as a requirement in all urban water development projects, there

Table 3. Water price variation in selected developing cities.

Water prices, US cents per 1000 litres

Water source Lima Kampala Bandung Dar es Salam

Utility provision 0.5 – 0.5 –


Kiosk – 1.0 – –
Water truck 2.3 – – –
Water vendor – 8.7 3.5 –
Handcarts – – – 1.6
Stand pipes – – – 5.0
Do investments and policy interventions reach the poorest of the poor? 229

should be some provision for an open access water point to provide for the needs of those who
otherwise may slip through the net.
One way in which this the problem of subsidization can be considered on a much larger
scale is in the case of large water projects. In many countries, both irrigation and hydropower
schemes often benefit from high degrees of government subsidy, although often these are not
made clear. In such cases, the real costs associated with water consumption for these purposes
are to some degree unaccounted-for, and it may be that the beneficiaries are not really those
in the greatest need. Examples of this can be seen in numerous countries where dams and
reservoirs which are built are actually used largely by private commercial farmers, or power
companies, rather than the poorest of the poor. One way of overcoming these problems would
be through the development of more transparent systems of subsidization, more effective
prioritization of investments, and the development of better identification of those who are in
the greatest need.
In the overcrowded cities of the world, millions of people may be within a short distance of
a water supply point, but in practice, this may not be available to them. It could be that they do
not have legal access to this water, or that they are prevented from its use by caste or other social
constraints. In addition, in most urban settings, there is a huge proportion of household time
spent queuing to get access to the water points, and single points may be shared by thousands
of households. This may be simply due to the huge number involved, or due to low pressure
reaching outlying shantytowns where so many have to make their homes. This is one of the
reasons why some argue that the data generated by the Joint Monitoring Programme (WHO-
UNICEF, 2000) does not provide an accurate picture of water deprivation, as that information
is simply based on physical distance from a water source. The additional time spent in queuing
to have access to water points represents another example of the unaccounted-for aspects of
water management.
Today, in many cases, poverty is increased by ecosystem degradation, and in its turn,
poverty may create more degradation (Kaufmann, 1995). At the same time, when people are
empowered to manage their own resources, this can be extremely beneficial, and can bring
about a real change in the way resources are used for different purposes. There is little doubt
that there is much need for more effective monitoring of the links between water and poverty,
and it is important to identify those places where the need for improvement is greatest, thus
enabling the prioritization of action. By identifying and tracking the physical, economic and
social drivers which link water and poverty, decision-makers will be able to identify who needs
water, where, thus enabling action to be taken on the ground to meet these pressing needs.

10 CONCLUSION

This chapter has tried to argue that while policy interventions and investments have a sig-
nificant role to play in poverty alleviation, they often fail to account for the most vulnerable
groups of the poor. It is recognized in most countries that the activities of a sizeable sector of
the population are not included in national income statistics. These are the unaccounted-for
millions: often very poor people who participate in activities which are not formally recog-
nized and recorded. People classed as the rural poor are often in this category, since their
economic activities are largely not accounted for. The majority of rural people are likely to
generate their own food, consuming it directly, without recourse to a market. In addition, they
230 C.A. Sullivan

may gather products from nature such as fuelwood, along with forest foods and medicines. In
societies where capitals are tied up in livestock, people are reluctant to reveal the extent of
their livestock holdings, and it may be the case that the ecological costs associated with their
activities are also unaccounted-for.
There is no doubt that there is a vital role to be played by investments and policy interven-
tions in the water sector. Nevertheless, overcoming inequities in water allocation is one of the
objectives of improving stakeholder participation and better water governance, and there is a
clear need for more inclusive and accurate assessments of access to water, and the way it is
used. While agriculture in general uses around 70% of all water used for human consumption,
and the level of domestic water needed to meet all human health needs is less than 5% of total
withdrawals, it is clear that there could be a small amount of water taken from agriculture
(through improvements in water use efficiency), to provide for better domestic provision.
While this suggestion may be said to be simplistic, there is no doubt that in a number of
countries, great improvements could be made to the quality of life of millions of people if
more account was taken of what was unaccounted-for, and if more attention was paid to
building the political will necessary to achieve this.

REFERENCES

Beckerman, W. (1995). Small is stupid: Blowing the whistle on the Greens. Duckworth, London, UK.
Briscoe, J. (1996). Water as an economic good. Paper presented at the World Congress of International
Commission on Irrigation and Drainage. Cairo, Egypt.
Costanza, R.; d’Arge, R.; de Groot, R.; Farber, S.; Grasso, M.; Hannon, B.; Limburg, K.; Naeem, S.;
O’Neill, R.V.; Paruelo, J.; Raskinet, R.G.; Sutton, P. & van den Belt, M. (1997). The value of the
world’s ecosystem services and natural capital. Nature, 387: 253–260.
Daly, H. (1999). Ecological Economics and the Ecology of Economics. Edward Elgar, Cheltenham, UK.
Desai, M. (1995). Poverty, Famine and Economic Development. Edward Elgar, Aldershot, UK.
Dreze, J.; Samson, M. & Singh, S. (1997). The Dam and the Nation: Displacement and Resettlement in
the Narmada Valley. Oxford University Press, New Delhi & Oxford.
El Serafy, S. (1991). The environment as capital. In: R. Costanza (ed.), Ecological Economics: the
Science and Management of Sustainability. Columbia University Press, New York, USA.
Falkenmark, M. (1990). Rapid population growth and water scarcity: the predicament of tomorrow’s
Africa. In: K. Davis & M.S. Bernstam (eds.), Resources, environment and population. Supplement to
Population and Development Review, 16: 81–94.
Farber, S. & Costanza, R. (1987). The economic value of wetland systems. Journal of Environmental
Management, 24: 41–51.
Faucheux, S. & O’Connor, M. (1998). Valuation for Sustainable Development. Edward Elgar,
Cheltenham, UK.
Friend, A.M. (1993). Feasibility of environmental and resource accounting in developing countries.
Environmental Accounting: A Review of the Current Debate. UNEP, Environmental Economics Series,
8. Nairobi, Kenya.
Gleick, P. (2000). The World’s Water 2000–2001. Island Press, London, UK.
Gordon, D. & Spicker, D. (eds.) (1999). The International Glossary on Poverty. Zed Books, London,
UK.
Hannon, B. (1991). Accounting in ecological systems. In: R. Costanza (ed.), Ecological Economics: the
Science and Management of Sustainability. Columbia University Press, New York, USA.
Hueting, R.; Reijnders, B.; de Boer, B.; Lambooy, J. & Jansen, H. (1998). The concept of environmental
function and its valuation. Ecological Economics, 25: 31–37.
Jacobs, M. (1994). The limits to neoclassicism. Towards an institutional environmental economics. In:
M. Redclift & T. Benton (eds.), Social Theory and the Global Environment. Routledge, London, UK.
Do investments and policy interventions reach the poorest of the poor? 231

Jacobs, M. (1997). Deliberative Democracy. In J. Foster (ed.), Valuing Nature. Ethics, Economics and
the Environment. Routledge, London, UK.
Kaufmann, R.K. (1995). The economic multiplier of environmental life support: can capital substitute
for a degraded environment? Ecological Economics, 12: 67–79.
Lipton, M. (1988). The Poor and the Poorest. World Bank, Washington, D.C., USA.
Lumby, S. (1979). Investment Appraisal and Financing Decisions. Chapman and Hall, London, UK.
Malseed, J. (1990). Bread without dough: understanding food poverty. Horton, Bradford, UK.
Meigh, J.R. & Sullivan, C.A. (in press). The Impact of Climate Variations on Water Resources: an
Index-Based Approach to Assess Human Vulnerability. Climate Change (Special Issue).
Opschoor, J.B. (1998). The value of ecosystem services: whose values? Ecological Economics, 25:
41–43.
Orshansky, M. (1969). How poverty is measured. Monthly Labor Review, 92(2): 37–41.
Postel, S.L. (1998). Water for food production: will there be enough in 2025? Biosciences, 28: 629–637.
Raskin, P.; Gleick, P.; Kirshen, P.; Pontius, G. & Strezepek, K. (1997). Water Futures: Assessment
of Long-Range Patterns and Problems. In: Comprehensive Assessment of Freshwater Resources
of the World (background document for Chapter 3). Stockholm Environment Institute, Boston,
Massachusetts, USA.
Rosenhouse, M. (1989). Identifying the poor: is headship a useful concept? LSMS Working Paper, 58.
World Bank, Washington, D.C., USA.
Scoones, I. (1998). Sustainable Rural Livelihoods: a Framework for Analysis. IDS Working Paper, 72.
IDS, Brighton, UK.
Sen, A. (1995). Mortality as an indicator of economic success and failure. Discussion paper, 66. London
School of Economics and Political Science, UK.
Simon, J. & Khan, H. (eds.) (1984). The Resourceful Earth. Blackwell, Oxford, UK.
Spash, C.L. (1998). Environmental values and wetland ecosystems: CVM, ethics and attitudes. In: Social
processes for environmental valuation. The VALSE project final report. Versailles, France.
Sullivan, C.A. (2002). Calculating a Water Poverty Index. World Development, 30: 1195–1210.
Sullivan, C.A. & Meigh, J.R. (2003). The Water Poverty Index: its role in the context of poverty
alleviation. Water Policy, 5: 5. October.
Sullivan, C.A.; Meigh, J.R. & Acreman, M.C. (2002). Scoping Study on the Identification of Hot Spots –
Areas of high vulnerability to climatic variability and change identified using a Climate Vulnerability
Index. Report to Dialogue on Water and Climate, Centre for Ecology & Hydrology, Wallingford, UK.
Sullivan, C.A.; Meigh, J.R.; Giacomello, A.M.; Fediw, T.; Lawrence, P.; Samad, M.; Mlote, S.;
Hutton, C.; Allan, J.A.; Schulze, R.E.; Dlamini, D.J.M.; Cosgrove, W.; Delli Priscoli, J.; Gleick, P.;
Smout, I.; Cobbing, J.; Calow, R.; Hunt, C.; Hussain, A.; Acreman, M.C.; King, J.; Malomo, S.;
Tate, E.L.; O’Regan, D.; Milner, S. & Steyl, I. (2003). The Water Poverty Index: development and
application at the community scale. Natural Resources Forum, 27: 1–11.
UNCED (1992). Report of the United Nations Conference on Environment and Development. United
Nations, New York, USA.
UNDP (2000). Overcoming Human Poverty. United Nations Development Programme, NewYork, USA.
UNESCO (United Nations Educational Scientific and Cultural Organization) (2003). World Water
Development Report. UNESCO, Paris, France.
Uvin, P. (1994). The International Organisation of Hunger. Kegan Paul, London, UK.
van der Gaag, J. (1988). Confronting Poverty in Developing Countries: Definitions, Information and
Policies. LSMS Working Paper, 48. World Bank, Washington, D.C., USA.
WHO (1992). Health for All Database. World Health Organization, Geneva, Switzerland.
WHO-UNICEF (2000). Global Water Supply and Sanitation Assessment 2000. Joint Monitoring
Programme for Water Supply and Sanitation. World Health Organization and UNICEF, Geneva,
Switzerland.
World Bank (1992). World Development Report. Oxford University Press, New York, USA.
World Bank (1996). Poverty Monitoring and Analysis Systems. World Bank, Washington, D.C., USA.
World Bank (1998). Standardised Welfare Indicators. World Bank, Washington, D.C., USA.
VII

Water and Nature


CHAPTER 14

Water and nature. The berth of life

F. García Novo
Estación de Ecología Acuática, University of Seville, Spain

F. García Bouzas
Junta de Andalucía, Seville, Spain

ABSTRACT: Life is presented as an historic phenomenon arising on the planet in close connection
to water environments and the water cycle. Early stages of life on Earth are underlined, pointing to some
relevant roles of water in cell physiology. The breaking up of water during photosynthesis is focused as
a far reaching step toward the development of life.
The importance of water in energy exchange and circulation of the atmosphere is discussed and
climates are characterised in terms of water availability, regime and temperature. The role of water in the
alteration of minerals and in transport processes of continents are discussed in terms of geomorphology
and soil evolution under each climatic type. As a summary, a classification of environments is presented,
showing how living organisms have succeeded in coping with varied conditions during the evolutionary
process.
Communities are introduced as the ecological basis of life in ecological systems presenting a few
examples of regulatory mechanisms involving water functions. Wetlands are presented in some detail
and the relationships of basic ecological traits such as community structure, diversity and productivity
in relation to water availability are discussed.
The advent of cultural man started a cascade of ecological phenomena that impaired the primeval
functioning of ecosystems. Environmental consequences of river impoundment, water diversion, irriga-
tion, and drainage of wetlands are dealt with. Eutrophication, contamination and salination of natural
waters are discussed. Exploitation of aquatic organisms such as fish and the overall effects of distur-
bance on biodiversity are also shown. The introduction of aquatic species to new biological areas and
the side effects of introductions on natural systems, rivers and cultivated areas and the invasion of urban
aquatic environments by wild organisms is presented in this perspective of human induced biodiversity
changes.
Some examples of subtle human alterations of aquatic systems are presented suggesting new technical
developments were needed to attain sustainability development policies.

Keywords: Water cycle, ecology of water, impacts on natural waters, water life, wetlands, water
cultures

1 LIFE ORIGINS

The origin of Life on our planet is still uncertain. The earliest evidence dates back some 3860
million years (Mojzsis et al., 1996), this evidence consisting of isotopically light carbonaceous
inclusions within grains of apatite (basic calcium phosphate), only known to build up within
living cells. The deposits belong to a sedimentary formation.
236 F. García Novo & F. García Bouzas

Early fossils were aquatic organisms. Terrestrial forms appeared much later in the timeline.
Life access to continents was mediated through aquatic forms invading estuaries, swamps
and areas with plentiful water. The expansion of living organisms from aquatic to aerial
environments allowed for the colonisation of continents and was followed by a large increase
in biomass and species diversity.
All known life forms basically are tight arrangements of (organic) vesicles containing water
solutions and dispersions, their composition being controlled by permeability and transport
functions of vesicles walls and membranes. The description of molecular mechanisms involved
in water transport across membrane pores represents a recent advance in biochemistry.
Life appeared in water environments and has evolved for most of its span as aquatic
organisms. Life performs in water.

2 ECOLOGY

For minute living forms, such as micro organisms, the diffusion process is sufficient to
exchange materials with the surrounding water environment. More efficient mechanisms
involve the incorporation of water droplets to the cell in the form of vesicles, letting organic
particles be exploited as trophic resources. The opening of fine ducts in the organism letting
water solutions flow through the body largely increased the surface available for exchange
of ions and gases or the delivery of particles. It had profound consequences in evolution:
circulatory systems allowed for the integration of a single organism composed of specialised
organs. The opening of outlets connecting ducts to the outer medium lies in the base of diver-
sification of digestion, excretion, respiration and reproduction. Multicellular organisms and
finally Metazoa, expanded as physiological water roles multiplied.
The appearance of contracting organelles led to pulsing structures in life forms, making
possible the flow of digestive tracts or circulatory systems. The formation of flagella let
organisms move, avoiding unfavourable environments and gave them the means to tap distant
resources. Dispersal to new areas or to patchy resources in aquatic media became feasible.
When sexual systems evolved, it was movement coupled to sensory mechanisms what allowed
individuals or complementary sexes to couple. Sessile organisms, unable to move, retained
a free mobile stage in water as gametes, larvae, or juveniles. Contracting cells arranged into
muscles added a longer range of movement for animals, letting them commute between distant
environments and to exploit resources, such as other organisms, on a more efficient scale.
The development of Life has exploited the water environment in a growing number of ways.
Water as source or sink for gases, ions, salts and particles has been mentioned. However,
diffusion flow proceeds slowly in still water as opposed to turbulent diffusion. Turbulence, or
energy dissipation inducing it, has been exploited by some life forms which adapted to the
littoral, to strong current sites, or thermal vents where turbulent diffusion occurs. At every
scale seas show currents that have been integrated by many organisms in their life cycles: from
tide or daily local currents to oceanic giro, all scales have been encountered and exploited
as an energy source (for transport), by one group of organisms or other. And evolutionary
sequences often document in organisms the appearance of convergent body forms and organs
imposed by water properties.
For displacement through water, a fusiform body provided with a round head and a thin tail
is the more efficient morphology. Propulsion organs should be in the rear. Fish-like forms are
Water and nature. The berth of life 237

shared by the different lineages of fish, amphibians, and cetaceans among mammals. Other
fast swimmers such as penguins, sea lions, and seals, have heads with big jaws or beaks and
palmed legs instead a long tail, altering their silhouettes, but the body shape sticks to the
fusiform pattern imposed by water viscosity.
As opposed to this pattern, planktonic species may evolve to bigger sizes and longer life
cycles. Size implies, according to Stokes’ Law, a faster precipitation. Vertical movements in
the aquatic column may displace the organism from the favourable environment. To impair
precipitation many plankton species possess chetae, appendages or specialised ramified organs
with a large surface.
Turbulence favours diffusion in the aquatic environment supplying sessile organisms with
nutrients, oxygen, and organic particles. The easy access to supplies cannot be separated from
the transport of mineral particles abrading epidermis or clogging ducts and respiratory organs.
Intense turbulence only occurs in certain locations where large energies are released to the
water such as the littoral where waves break or in steep river channels. In each case the energy
stored as wave oscillatory motion or the potential energy of the river flow is released, turning
into kinetic energy of the water.
A few morphological patterns have been adopted by those organisms surviving in turbulent
waters. For plants that demand light intercepting surfaces, laminar extensions shape into nar-
row bands forming belts, ribbons, narrow flat branches and threads. Brown algae (Feophyta)
living in littoral waters often exhibit a thallus with a strong anchoring organ, a stem and a
leaf-like expansion In Laminariales that may be laciniate such as Himantalia, or Chorda, dig-
itated as in many Laminaria, flat or thread-like branches as in Cistoserira or Fucus species. It
has been documented for Laminaria that the tensile shear induced by wave action elongates
the stem of the plant generating heat and favouring metabolic activity. Red algae (Rodophyta)
exhibit similar morphological patterns standing better water turbulence. It is worth noting that
many aquatic vascular plants from flowing waters adopt comparable morphologies with roots
and rhizomes fixing them to the substrate and strong laciniate leaves standing turbulence.
Zostera and Posidonia live on shallow marine waters; species of Ceratophyllum, Myriophyl-
lum, Potamogeton and many others are present in rivers. An extreme adaptation to intense
water turbulence is exhibited by species from Podostemonaceae, a tropical family living in
cascades, also presenting laciniate forms.
Animals do not need exposing a large body surface to radiation and confront water tur-
bulence with a strong skeleton and an adequate organ for fixation to a solid substrate. The
development of organs resistant to water shear and erosion, let animals exploit environments
with very high turbulence colonising cliff rock surfaces exposed to intense wave action. The
best hydrodynamic shape in response to water turbulence is a cone fixed to the ground. A
strong conical skeleton protects the animal soft organs inside. A foot secures the cone to the
substrate and an upper opening, often protected with a lid, keeps the communication to the
environment. Or else, the cone apex is turned outside and the base is attached to the sub-
strate. From early Palaeozoic era some coastal marine invertebrates from unrelated groups
have evolved to produce cone-shaped individuals: Archeocyatids (Porifera), Brachiopoda or
Madreporaria, or the Mesozoic Rudista molluscs, or the contemporary limpets (Patella), are
a few examples of shape convergence in response to water turbulence.
Sessile animals build complex colonies to exploit a moderate turbulence: the numerous
coral-like forms where individual animals trap organic particles or minute organism brought
in by water currents. In some cases, Red algae Criptonemiales, may approach this pattern the
238 F. García Novo & F. García Bouzas

thallus presenting calcareous deposits and the algae assuming articulate segments (Corallina,
Halimeda), or shaping into crusts in a lichen like form (Lithothamnium) or in coral like patterns
(Porolithon oncodes).
Vertebrates originated in aquatic media terrestrial forms having their origin in the fish-
pulmonate fish-amphibian-reptile sequence. It is intriguing that some of the terrestrial lineages
turned back to continental waters or to the ocean to exploit trophic resources or to live more
or less permanently in aquatic media. The extinct reptiles Ichtiosaurs and Plesiosaurs, the
numerous marine and aquatic turtles; among mammals, the cetaceans, seals and manatees,
beaver, otters, capybara (Hydrochoerus capybara); among birds, penguins, albatross, and the
large number of waterfowl are but a few examples of the process. As an old trait of their
terrestrial lineage, these vertebrates use lungs to breathe atmospheric air and need surfacing
from time to time. Many groups also demand terrestrial ground for reproduction: nesting,
egg hutching, rearing, delivering or mating. Breeding colonies regulate offspring numbers,
avoiding overpopulation bursts in the species.
Morphology usually conforms to aquatic habitat restrictions, but in several cases, terrestrial
morphologies are capable of efficiently exploiting aquatic resources. The morphologies of
the marine iguana of Galápagos (Amblyrhynchus cristatus), the dippers (Cinclus cinclus, C.
mexicanus), or the osprey (Pandion haliaetus) do not suggest their ability to dive, forage or
catch underwater. Brown and black bears (Ursus arctos, U. americanus) are great salmon
catchers in fast rivers. Examples are plentiful to show how water surfaces are crossed by some
species to exploit resources on the other side. These examples belong to this category and
water fowl, at large, is also included.
In a few cases some aquatic animals exploit terrestrial organisms: fish species preying on
insects flying above the water surface throwing a jet of water (Arawana) or reaching for the
fruits in branches of trees flooded in the high water, várzea, of the Amazon basin (Colossoma
macropomum). Some fish, such as the European eel (Anguilla anguilla) leave the water and
creep through meadows over long distances, behaving like snakes. Perioftalmus is a remarkable
fish from mangroves that abandons water to run, with the help of fore fins, on the muddy
shore searching for food. Large aquatic predators such as alligators and crocodiles often
attack terrestrial animals approaching the wetland or living in it. Piranha fish (Serrasalmus,
Pygocentrus), are capable of vicious collective attacks to terrestrial vertebrates venturing into
their waters.
The distinction between aquatic and terrestrial environments is blurred after intense rains
or under wet climates. Soil surface easily form pockets of water that may last for a few hours
or days, long enough for a temporary aquatic ecosystem, to develop. Anphipods are a group of
crustaceans that often dominate temporary ponds in mid latitudes. In tropical climates some
plant groups, notably Bromeliaceae, store rainwater among the sheath of upper leaves. These
minute water bodies present a distinct planktonic composition according to light intensity, plant
size and the bromeliad species involved. Sarracenia and Nepentes are genera of carnivorous
plants with modified leaves that store water. The decomposition of insects falling into the
water trap releases some nutrients usable for the plant. But the trap offers a durable water body
that allows for the development of amphibian toads. Bromelias, nepentes, and many other
plants serve as temporary nurseries for toads. Time and again water plays multiple ecological
roles.
Water in the soil profile plays some important roles in the processes: rainfall infiltration
through the soil dissolves minerals, bringing their elements to the soil solution that in turn
Water and nature. The berth of life 239

behaves like a chemostat in a dynamic equilibrium adjusting its concentration of ions, salts
and gases to the reactive sites of soil minerals, and to local hydraulic pressure. Vertical water
flow redistributes the elements from one soil horizon to the next. The slow water circulation
to deeper levels completely modifies lithosphere composition.
Water dissolution also creates pores and ducts in the rock matrix. Water movement in films
through soil structures carries away particles. The action of ice or the swelling and shrinking of
some clay minerals and organic matter deposits, according to their water content, also favours
mechanical movements in the soil all caused by water action.
With the expansion of plants to the terrestrial media, the water environment was restricted
to the soil matrix where the finer roots reach the water films among edaphic particles. Extrac-
tion of nutrients from the soil is largely mediated by the soil solution and is achieved by a
concentration process carried out by permeases, implying (metabolic) energy expenditure.
The strongly negative water potential in the atmosphere, as compared to that of organ-
isms, menaces terrestrial life forms with desiccation. If impermeable coats or thick epidermis
developed to prevent water losses, the abstraction of O2 or CO2 from the atmosphere was
impaired. Free communication with the atmosphere implied the occurrence of transpiration
for terrestrial organisms. Water economy became a key trait of terrestrial organisms exhibiting
powerful absorption organs, a stringent regulation of water losses, water recovery processes
and an efficient use of water in their metabolism. Water circulation and water loss opened new
evolutionary arenas to terrestrial plants and animals.
Vascular plants have developed the xylem, a vascular system made of bundles of fine
pipes with rigid walls capable of withstanding strong pressures. Vascular plant leaves couple
negative water tension in leaf parenchyma to water columns contained in xylem tissue bringing
negative pressures to root level. Soil water is easily absorbed by the plant, accomplishing two
relevant functions: to compensate for transpiration losses and to bring fresh soil solution to the
roots, supplying them with fresh nutrients. It should be noted that the energy needed for root
water absorption and transport to leaves is largely supplied by the atmosphere as a gradient
of water potential from air to soil. Together with photosynthesis, this mechanism represents a
net input of energy to plant metabolism.
For terrestrial animals, desiccation has selected peculiar ecological traits such as the imper-
meable cuticles of many insects, the loss of flying wings (i.e. Tenebrionidae beetles), the
recovery of water from faeces and urine (i.e. reptiles of very dry habitats), the storage of water
inside the body (such as in the toad Pelobates cultripes), and others.
Many diurnal animals benefit from sun radiation, warming up to increase activity. It is a well
known condition for lizards and many other reptiles, including aquatic turtles, alligators and
crocodiles. Birds and mammals have evolved homothermal physiological mechanisms where
activity from muscles and other organs releases heat that is lost through thermal conductivity
through the skin and specialised organs. Transpiration can play an important role here for
homotherms as water evaporation from skin absorbs a large heat load letting the body regulate
temperature.
Soil horizons are somehow protected from desiccation; consequently, edaphic organisms
often benefit from moist environments and small surfaces of water films. On a minute scale,
soil structure with long connected pores among granules of minerals, small rock fragments
and organic matter particles, can compare to a network of tunnels with water pockets.
The soil surface receives fresh organic matter from vegetation, less so from animals. Mineral
particles are brought by wind or incorporated into rainfall together with dissolved materials.
240 F. García Novo & F. García Bouzas

Organic matter contains the energy supply for decomposers, a large fraction of soil organisms
that help structure the soil by breaking down the organic compounds, finally producing humic
and fulvic acids and releasing some of the mineral ions to the soil solution. Clay minerals often
arise from secondary formation within the soil profile. The large exchange capacity of humus
favours the storing of nutrients in the soil profile. Water storage (field capacity) also benefits
from the accumulation of organic matter, fine particles and clay minerals in this profile.
Vascular plants are favoured by the expanded soil retention of nutrients and larger water
reserves, showing a larger primary productivity that, in turn, sustains a higher secondary
productivity. At the same time vegetation intercepts some of the precipitation before reaching
the surface and pumps out water reserves in transpiration, reducing soil moisture.
The building up of a soil horizon reduces runoff and stores water in the profile. It is through
these mechanisms that terrestrial vegetation plays a significant role in the soil water balance
and the partitioning of precipitation among evapotranspiration, infiltration/aquifer recharge
and runoff. Runoff depends on precipitation/infiltration rate, surface slope and vegetation
development. Living communities actively control transpiration to the atmosphere, and in
some cases induce higher precipitation by means of interception of low clouds crossing through
the canopies in cloud forests. In wet tropical climates it is remarkable the variety of plant
epiphytes growing on leaves, stems or trunks of perennial species, taking advantage of air
moisture replenished by cloud circulation and precipitation. In some vegetation types such as
Laurisilva it is the aerial structure of vegetation that favours precipitation inside the forest as
low clouds move through it.
Should the forest be suppressed, precipitation would be much reduced, completely alter-
ing infiltration and runoff. Soil horizons would lose most of their organic matter and their
field and exchange capacities would be much reduced. Logging operations under the above
circumstances often result in irretrievable devastation to ecosystems.
On Hierro Island (Canary Islands, Spain), it was the mighty Garoe tree (Ocotea foetens)
that supplied water to meet the islanders’ needs before Europeans settle the island. A hurricane
uprooted the tree by 1610, and the population, deprived of water, fled from the island.
A few animal species have also developed some mechanisms to tape water from atmosphere.
The fog basking beetle (Onymacris unguicularis) from Namibia desert climbs to dune crests
where it adopts a head down stance with the dorsal surface exposed to the fog bearing winds.
Droplets slip towards the head and the beetle can gain a 12% weight in water in a single night
(Cloudsey-Thompson, 1991).

3 CLIMATE AND THE WATER CYCLE

The water mass of the planet has been estimated to be 1600 million km3 , and 97% belongs to
the oceans including water and ice. Annual evaporation from ocean represents 448,000 km3
and another 72,000 km3 come from the continents1 . 100 km3 daily water flows into the sea,
37,000 km3 annually. Precipitation on continents averages 730 mm/yr supplying vegeta-
tion, refilling wetlands, keeping drainage network flowing and favouring circulation of
biogeochemical cycles. Water scarcity in dry lands or water in ice rigorously limit life
development also bringing rock alteration and sediment transport to a reduced state.

1
Water balance figures change slightly with published source.
Water and nature. The berth of life 241

Water plays important roles in the circulation of energy through the atmosphere. The major
energetic circulation takes place as radiation. Infrared Radiation (IR) exchanges between the
lower troposphere, and the ocean and continent surfaces are about four times as important as
Photosynthetic Active Radiation (PAR) received from the sun at the continent surface level. At
IR wave lengths water vapour exhibits important absorption bands (4–7 and over 18 micron).
It intercepts a fraction of solar radiation and more strongly the Earth irradiation fluxes that
largely correspond to a 4–20 micron interval. Cloud water droplets are efficient absorbers but
also serve to reflect radiation at both visible and invisible wave lengths. From outer space, our
Blue Planet exhibits strong white patterns of cloud formation against a blue background of
oceans and shades of vegetation blurred by atmosphere.
Energy flows in the troposphere are also associated with mass air displacements. The
radiative energy absorbed by soil and ocean surfaces, heats these outer layers that in turn
transfer some of the energy to the base of air column raising its temperature (sensitive heat).
Expanded air masses become unstable and tend to rise in the troposphere until they find a new
equilibrium point inducing convective and advective air movements. A second mechanism for
energy transfer (about 5 times more) is water evaporation from ocean or continent surface
to the air. The amount of water involved approaches 36 mg/m2 /s and the associated energy
flux is about 82 W/m2 , averaging the flow for the planet surface. Plants behave as evaporating
surfaces or intermediaries between atmosphere and soil in the transpiration process.
Average residence time for water in the atmosphere is short: 3 to 4 days, but the relevant
process is that water enters the atmosphere as vapour and leaves it as water or ice in rain, hail
or snow. The energy difference per gram of water mass in the cycle varies between 600 and
700 cal. The water cycle thus appears as a powerful energy transfer mechanism from oceans
and continents to atmosphere. Energy budget is dissipated through winds that mix atmospheric
layers favouring the vertical transfer of gases and air colloids and fine particles. Horizontal
winds transfer air masses from source areas, redistributing water and energy, lowering the
temperature and precipitation at low latitudes and increasing it at high latitudes. Some wind
energy is also transferred back to the ocean inducing sea currents and waves.
Temperature and precipitation patterns at surface level largely define climate. Mineral
alteration of rocks and life development are both favoured by high temperatures and water
availability. Cool climates, where water remains ice or in arid climates where water is scarce
and easily evaporates, show limited mineral alteration and a limitation of living forms as
well as low productivity for continents. For oceans, life is loosely associated with climate
because temperature plays a modest role and precipitation has little or no importance. Instead,
it is PAR radiation, having to do with latitude, and nutrient availability, that play major
roles.

3.1 Continents: the slope/river continuum


Rainfall is a key character to the continents of our planet. Every year precipitation brings
to the continent surface 108,000 km3 of water with low mineral contents. This represents a
dual energy storage: water is elevated against the gravitational force storing potential energy.
Average continent height is 823 m, although altitude precipitation is uneven. A second energy
source is chemical water potential that will lower as more substances are dissolved along the
lithosphere course.
If precipitation is snow, and no melting occurs in the area, water effects are limited to rock
abrasion and transport, with limited chemical alteration, generating a peculiar morphology
242 F. García Novo & F. García Bouzas

on exposed rocks and deposits. When ice finally melts fluvial processes reshape glacial geo-
morphology. When glaciers reach the sea they sometimes form an ice field that breaks up into
icebergs. Icebergs represent an ice transport mechanism that carries out rock fragments to some
distances over the sea, following the currents that drive ice floes away from their origin. The bot-
tom of Labrador Channel is littered with drifting materials and boulders carried out by ice foes.
The action of water on rock surface depends on the mineral composition. When a single
or few mineral species dominate (as in limestone, dolomite, gypsum, halite and carnalite or
quartzite), the rock mass dissolves and is carried out in solution by water. Joints, fractures,
strata discontinuities, easy water circulation inside rock openings, pipes, vertical chimneys,
open a network of ducts and cavities inside the mass. Limestone formations are outstanding
for the development of caves, ducts, inner lakes and the formation of deposits inside then. In
a similar way, circulating water opens ducts and cavities inside the ice mass in glaciers.
When parent rock is composed of several minerals, alteration rate depends on temperature
and water availability. Easily altered minerals yield ions, salts, or lattice arrangements where
some ions or elements are substituted for others. The rock and the circulating solution interact
with one another for as long as water moves through it in a large series of processes involving
exchange, solution, substitution, oxidation, hydration, carbonation, precipitation, eventually
erosion or deposition.
Under different climates (temperature, water precipitation), rock alteration proceeds dif-
ferently, with a different array of sensitive vs. resistant minerals under each condition. Thus
the water solution will carry away a variant balance of elements and a new assemblage of
unaltered resistates will stay longer on the rock surface.
The partition of rainfall water between infiltration and runoff depends on the surface bal-
ance: where rainfall excess over infiltration feeds runoff flow. The sheet of surface running
water will exhibit a low residence time, moving downslope with little time to develop a chem-
ical interaction with surface minerals. On the other hand, as water concentrates on a few lines
of flow, velocity will rise and turbulence appears, permitting the turbulent transport of soil
particles. As runoff channels concentrate higher debits, water velocity increases and channel
slope, width and depth adjust to water flow in a process that will extend to the whole watershed.
Runoff water initially moves resistate minerals as sands and rock fragments, as gravels.
Other materials from rock alteration are also available to transport, such as clay particles, salts
and ions. Organic compounds and particles, including minute organisms, will also be carried
downslope by water turning potential to kinetic energy.
Rock materials under adequate conditions move downward by mass wasting processes
such as solifluxion, creeping, slumping and others. The sheet of loose materials is regolith.
Its rate of displacement depends on the slope angle, easing as it decreases. The increase of
water content reduces viscosity of regolith, favouring creeping. In solifluxion it is through the
volume changes of water in thawing-melting cycles that makes regolith progress. It becomes
even more important for slumping where a regolith sheet, in the form of a large slate, slips on
a surface, often lubricated by the presence of water. In a mudflow it is the regolith saturated
in water that flows downslope forming a semi-liquid current of high density and destructive
effects that will stop its progression when reaching a flat area, where no more potential energy
conversion is possible.
Slopes are the key morphological elements of landscapes and it is in the geological processes
building them that surface water is present at every level. Alteration and erosion prevail at
the top of the slope on subhorizontal to moderately steep segments. Transport dominates the
intermediate steeper segments. Sedimentation will occur at the lower part of the slope that
Water and nature. The berth of life 243

tends, again, to be subhorizontal and where a channel usually drains the excess water feeding
the drainage network.
Infiltrated waters will develop soil structure through a sequential exchange of materials
with soil horizons from accumulation of organic matter and elution in the upper (A) horizon,
accumulation in the underlying (B) horizon and fresh rock surfaces where alteration was
starting at the deeper (C) horizon.
Water shapes the morphology and dominates mineral alteration and transport according to
climate. In each case a sequence of different habitats is generated, to which organisms adjust.
In surviving, communities often regulate the habitat.
The above description underlines the continuity of processes involving continental waters,
both in time and space. Water performs as a ubiquitous media renewing itself in volume and
composition and connecting biosphere systems. Drainage channels, rivers and lakes behave
as a connection network for living organisms. But in addition, water acts as a barrier limiting
energy exchange, mineral cycling, gas diffusion and organism displacement. One such a case
is ice, since under low temperatures runoff and infiltration come to a stand still. Snow or
ice covers are ecosystem shields preventing animal displacement or exploitation of resources
(such as foraging) across them.
The lower density of ice leaves it floating on water, severely restricting exchanges of the
water body, or the sea, with atmosphere. Stratification is an important outcome of water
density that depends on suspended sediments load, salt contents and temperature. For pure
water and at moderate pressures maximum densities are attained at 4◦ C. An energy input is
needed to raise to the surface the lower and denser layers close to 4◦ C, above the upper and
lighter ones. In consequence, stratification is a stable arrangement of the water column for as
long as densities remain unchanged and no energy input upturns the column.
Surface heating for incoming radiation reduces water density in summer, reinforcing strat-
ification. Cooling by irradiation, evaporation or exchange with a cool air mass, brings the
column to a point of instability when surface densities exceed the value of deeper levels. Cool
waters sink bringing to the surface warmer ones until the column reaches a new equilibrium
or the whole water mixes and the same temperature extends from surface to bottom. If cooling
proceeds, water masses approaching 4◦ C attain the highest densities (other variables being
equal) and collect at the bottom of the lake basin. Lower temperatures induce less dense water
bodies that now pile above the densest 4◦ C waters until the surface is finally frozen.
Stratification of the column whether caused by temperature or solid content differences con-
fines water environments to sub-horizontal layers. Of these it is only the superficial one that
receives light and exchanges gases with the atmosphere. Lower layers, deprived of light, are
inadequate for photosynthetic organisms, and O2 production ceases, but decomposition pro-
ceeds, often resulting in anoxic conditions and a higher nutrient content. Stratification confine-
ment induces chemical and biological differentiation of water environments leaving an upper
illuminated epilimnion where primary productivity persists. The upper layer behaves as the
collector of metabolic energy for the whole column. It is the input of organic matter synthesized
at the euphotic section that maintains all trophic levels in the dark section of the column.
Minute scale stratification operates in ponds. The seas and the oceans also exhibit the same
pattern of stratification where living forms gather at two levels: close to the surface where
light supplies the energy and on the bottom surface where the shower of organic particles
provides energy to the biological community.
For as long as water was available life thrives. But it is also water that separates or connects,
that makes terrestrial media favourable or adverse for each species.
244 F. García Novo & F. García Bouzas

4 COMMUNITY LEVELS

The ecological effects of water are not mediated by a single species, however important it was
for biomass or for cover. It is the community as the ecological assemblage of plants, animals,
and micro organisms, interacting above and below the soil surface, the water column and the
bottom surface that regulates matter, energy and, to some degree, structure. The outcome of
regulation is most varied, as shown with a few examples.
In temperate arid areas with scarce and irregular precipitations, such as the Negev Desert
in East Mediterranean, soil surface is largely covered by a crust of lichens and terrestrial
algae that absorbs some water when rainfall occurs but prevents infiltration favouring runoff
that concentrates in scattered points where perennials grow, thus producing patchy vegetation.
Should the lichen crust be disturbed, runoff would not concentrate and perennials fade away.
In mid latitudes, the regions dominated by high atmospheric pressure often show deserts
(Belnap & Lange, 2001). They exhibit scanty perennial vegetation, crusts of lichens or algae
and buried seed banks of annual species. These ephemeral plants quickly germinate after a
rainfall and produce seeds in just a few weeks of growth for as long as soil moisture lasts.
In cool and moist areas mineralization of organic matter proceeds slowly, building up
strong organic horizons. Primary productivity of vegetation remains low, but formation of
peat horizons frequently occur on flat areas, forming mires with water logged soils. Peat
levels rise above the soil level, forming large heaps of ombrogenic peat fed only by rainwater
and some dust brought in by winds. In boreal regions bogs, mires and peat deposits are
widespread, often displacing forest and grassland vegetation through water logging of soils.
Under wet tropical climates the two major factors enhancing productivity are favourable to
vegetation: temperatures are high and water supply is plentiful during the whole year or a large
part of it. The combination provides at community level, a plentiful primary productivity and
also larger productivities of consumers and decomposers. The accumulation of biomass per unit
surface in the form of timber, leaves and stems, epiphytes, algae, moss and lichens, soil fungi,
moulds, invertebrates or micro organisms is also high. The deposits of necromass as shed-off
organs, seeds and fruits, twigs and branches are impressive and pile up on the soil surface.
Per unit surface of ground, the fluxes of energy, water and minerals are high although they
exhibit a dynamic equilibrium through the operation of multiple regulatory loops based on
large biological diversities. There are many species present at each level; even the large trees
usually belong to several species that rise to tens of meters, with trees of all ages that can
reach the figure of 1000 per hectare. In a single tree may coexist several thousand invertebrate
species (Erwin, 1983).
This rich and diverse system tropical forest is sensitive to intervention and disturbances
because the uncoupling of regulatory processes easily unbalances it. Logging activities destroy
the structure, leaving many epiphytes and animals with a scarcity of ecological niches. The drop
in organic matter input in the soil does not impair the strong mineralization, which results in a
depletion of nutrients and a change in the surface soil horizon. Heavy rainfall regimes lead to
soil leaching. On steep slopes this may favour erosion and the onset of a vegetation dominated
by species of large leaves and fast to very fast growth and the presence of vines and climbers,
often making huge shields, hanging screens that profoundly alter structure and functioning of
the community. For partially unknown reasons, many animal species are particularly sensitive
to disturbance and community fragmentation suffices from their disappearance.
In temperate or warm climates where a dry season occurs, fires play a significant ecological
role. The Mediterranean type climate is such an example, occurring in the Mediterranean basin
Water and nature. The berth of life 245

and other mid latitude areas such as California, central Chile, West Australia and West South
Africa. In all these cases wildfires are naturally present, biomass and necromass accumulation
and nutrient deposits are lost in a recurrent way through fire. Loss of the vegetation structure
and the deposits of cinders on the soil surface ease runoff transport of nutrients when rainfall
reassumes after the dry season.
Wildfire occurrence is not limited to Mediterranean type climates: a continental climate
with a dry hot summer is prone to fires. Subtropical and tropical climates with a dry season
also produce natural wildfires. In a sense, it is water through rainfall patterns that controls
wildfire occurrence. The Mediterranean shrub vegetation of chemise, garriga, frygana, jaral,
apparently has gone one step further, integrating the fire event and the ensuing erosion process
of autumn rainfall into a mechanism of survival that prevents other plant communities from
invading its area (Herrera, 1992).

4.1 Wetlands: the productive waters


Shallow aquatic ecosystems on continents or coastal areas exhibit communities with a high
number of species, some in large populations. The reason, apparently, has to do with their
high primary productivity that maintains a wide base for trophic chains. Accumulation of
organic matter modifies nutrient balances, water regimes and the soil level of the wetland.
Fermentation of organic debris suffices to raise the temperature of their deposits, helping
some large reptiles such as alligator or caiman species to hatch their eggs. Wetland biomass,
dominated by photosynthetic organisms, includes phytoplankton in the water column and
vascular species rooting in the bottom and emerging to the surface. Many life forms may
coexist from trees such as mangroves, swamp cypress, tamarisks, poplars, to minute sedge
and rush species. Wetland plants often exhibiting leaves inside the water column such as
Utricularia or some Potamogeton species or spreading over the surface such as in Ranunculus.
Floating leaflets of Marsilea or round leaves of Nuphar and Ninphaea may almost cover water
surface, growing to an immense size of ca. 2 m across in Victoria regia leaves. Free floating
plants such as Eichornia crassipes, Trapa natans or Pistia stratiodes form a thick surface cover.
Smaller species also form almost continuous blankets such as species of gen. Salvinia, Lemma,
Ricciocarpus. Wolffia arhiza, a minute water lentil, with leaves 0.5–1 mm across, is the smallest
vascular plant. Some of the aquatic species have become aggressive invaders spreading to new
areas: Elodea canadensis in European channels; the water lilies Eichornia spp. in tropical
wetlands, and the contemporary invader in South Spain: Azolla caroliniana, the mosquito fern.
Animal groups benefit from the ample primary productivity in wetlands. Zooplankton and
macroinvertebrates that sustain vertebrate trophic chains of fish, amphibians, some reptiles
such as water turtles and crocodiles and many snake species live in tropical wetlands. Mammals
are present in small numbers with several rodents, bears, ungulates, manatees, a few large cats,
and others. The hippopotamus (Hyppopotamus amphibius) is the largest wetland mammal.
Wetlands offer a substantial habitat for many bird species exploiting primary productivity
(algal mats, fruits and seeds, tubers, leaves) or the net productivity of invertebrates, fishes and
amphibians. Large concentrations of birds forming breeding colonies in trees, shrubs, banks,
or amidst the emergent vegetation are a key character of many wetlands. Huge bird gatherings
numbering into the hundreds of thousands, possibly millions, generate an important effect on
vegetation and on the circulation of nutrients of the system.
Wetlands offer a breeding ground for many birds, fish and other living creatures that will
later spread to broader areas, possibly migrate to different regions. Other species will not breed
246 F. García Novo & F. García Bouzas

but come to the area to exploit the temporary bonanza of resources. Or endure a development
phase in their lifecycle such as the exchange between fluvial and marine environments or
the larval phase of many insect species. The wetlands keep unusual habitats where endemic
species survive and diverse communities have been maintained over long periods in spite of
climatic changes, thus preserving old ecological legacies.
Considering above arguments, the efficient removal of nutrients and sediments, and the
nursery effect on commercial species, coastal wetlands have been rated on an economic
basis, among the most important ecosystems to mankind in the comparative study of world
ecosystems carried out by Constanza et al. (1997).

5 MAN’S INFLUENCE

The advent of cultural man started a cascade of ecological phenomena that impaired the
primeval functioning of ecosystems. The industrial revolution that flourished mid 17th cen-
tury in Central Europe expanded environmental impacts beyond regional boundaries. It was
during the second half of the 20th century when global impacts on atmosphere composition,
ozonosphere, ocean fisheries, forest cover and biodiversity were first acknowledged. The early
years of the 21st century reinforced the scientific evidence for a global climate change that
also induced sea level variations. World water resources entered an era of incertitude.
Water availability severely limits biosphere productivity in warm and dry climates. The
historical dependence of agriculture and husbandry on precipitation was addressed by a rich
array of water technologies. This led to powerful water cultures emerging as actors of History
in the 4th Millennium BC. The early Empire development has been related to large scale
water management for irrigation (Wittfogel, 1957). In the vicinity of the Mediterranean basin,
Mesopotamia, the Nile Valley, the Fertile Crescent and the Arabian Peninsula agricultural
empires flourished prior to the 1st Millennium BC. This led to important cultural and social
developments, large population increases and depletion of local resources (fauna, forests,
mines) at an early phase.
Technology limitations prevented significant river impoundments but river diversion by
channels was feasible, greatly increasing crop production. Wetland drainage through channel
building turned swamps into agricultural lands. The combination of both strategies (such
as in the Tigris-Euphrates marshes in Mesopotamia, the Nile-Fayum in Egypt) represents
an early achievement for water technology. Another important step was the construction of
subterranean channels (quanat, foggara, mayra, viaje, galería), preventing evaporative losses
and permitting the exploitation of distant aquifers.
The combination of wells, vertical and horizontal water-wheels, channels, quanats and
drainage networks spread over the Mediterranean in successive waves, traversing the basin
from East to West. Significant advancements on water technologies were introduced under
the Roman Empire (approximately 3rd century to 4th century BC): long channels, aqueducts,
tunnels, lead pipes, water decanters, fountains, urban drainage systems (cloacae), sewers, field
drains (cuniculi). From the 7th century, Islam expansion in the Mediterranean introduced new
irrigation technologies from Mesopotamia and the Middle East including large water wheels
kept running by river flow, chain pumps, water distributors for field irrigation and gave new
impulse for quanat construction.
For centuries, available water technologies changed little but there was a steady improve-
ment in channel design, creating a fresh drainage network in Central Europe and England, and
Water and nature. The berth of life 247

finally developing internal transport systems. Dam building also improved, permitting some
water regulation for agriculture. The Dutch polder development coupled to water abstraction
by windmills permitted marsh reclamation for agriculture. European technology spread to
America after 15th century.
Water wheel mills were common under the Roman Empire for grain grinding, a tiresome
domestic task. They regained importance in Europe with the economic development occurring
ca. 1000. The small mills (1–2 kW power output per wheel) were used for grain grinding. But
step by step other energy demands were met by hydraulic wheels: metal processing, paper
production, wool cleaning and preparation, spinning, and other tasks. Tide mills were added
to river mills during medieval times. Water offered an adequate source of mechanical energy
for industrial purposes, fuelling the early steps of the Industrial Revolution. Turbines were
added early in the 19th century.
The advent of the steam engine in the last quarter of the 18th century completely altered
the picture: the Watt-Boulton engines were successful in every industrial branch and were
soon adapted to transport. The outcome had profound effects on industrial production, which
became cheaper and of a higher quality. New scales of production, new industrial processes
and a different type of human labour also emerged. From the environmental side, a few points
should be underlined.
The application of steam engines to pumps allowed the drainage of wetlands on an unpreced-
ented scale. It also permitted the extraction of water from rivers, providing water for industrial,
urban and agricultural uses. Mining was severely limited by water infiltration into deep gal-
leries. Pumping the excess water paved the way for deeper wells and again the steam engine
permitted the easy lift of ore, gear and man power. Intense mining resulted in widespread
washing of ore and the building up of slag heaps. Coal regions lost their rivers to contamin-
ation from coal dust. Coal and mineral sulphur mines released acid effluents from galleries
and tips, also degrading riverine ecosystems to a state beyond recognition: a flow of diluted
sulphuric acid rich in metal cations. The Rio Tinto River in Southwest Spain has suffered this
type of contamination for centuries and has turned to an almost abiotic river. The longstanding
contamination has suppressed original flora and fauna only surviving primitive chemosyn-
thetic life forms that thrive in the acidic waters and a heath (Erica andevalensis) growing on
the toxic slates left by Roman Empire miners.
The spread of railway networks was made possible by the steam engine. In turn it was this
easy land transport, combined with steamers on the high seas that favoured the distribution
of manufactured products, raw materials and the exploitation of minerals, forests, fisheries,
along with development of range and agriculture at rates never before attained. For Europe, the
second phase of the Industrial Revolution was a period of enormous industrial development,
urban growth, social change, and migratory movements. Other continents underwent a period
of invasion from European powers with large settlement of Europeans in mid-latitudes. The
vast human movement created new cities from scratch that quickly became large then important
trade and industrial centres. Buenos Aires and New York were old historical settlements in
America. But Chicago or Montevideo, small villages at the beginning of 19th century grew to
the million mark by 1900, surpassing Buenos Aires and approaching New York in population
during the following century. The Great Plains, to the West of the Great Lakes were entirely
transformed from forest and prairie to corn and cattle in less than a century: the Mississippi
basin was turned into a different entity.
From old historical cities large modern cities evolved, fed by migration, trade and indus-
try. Their vast populations were supplied with water from reservoirs or pumped from rivers.
248 F. García Novo & F. García Bouzas

Domestic waste water, industrial effluents and street drainage were collected and discharged
into neighbouring rivers. Certain micro-organisms survive easily in urban sewage, some tol-
erating the process of river self-purification, so as to reach the next water supply downriver,
affecting another population. Cholera and typhus outbreaks were commonplace. Parasitic
worms remained in the population through repeated contagion.
Aquatic flora and fauna endured an impact related to the ratio of wastewater discharge to
river flow. Large settlements such as Paris and London turned their rivers into stinking sewers
almost void of fish. Industrial regions such as the Ruhr destroyed the biota of their rivers. Indus-
trial ports contaminated coastal waters and estuaries. Water supplies became impossible and
many water resources such as fish or sea food, vanished. At present, the large urban and indus-
trial impact of sewage on continental and littoral waters has been controlled … in developed
countries. Polluting activities survive (or have been transferred) to less demanding countries
where politicians accept a relatively high degree of disturbance, usually because they lack funds
to tackle the problem. Water, in many forms, has entered our societies as a valuable asset.
Water has been the berth of Life, it paved the way for the evolution of life forms and
the building up of communities and ecosystems. For an enormous period of time, natural
regulation commanded the biosphere; it was only during Holocene, 10,000 years ago, that
man began to gain control. The last 100 years have witnessed overriding changes imposed by
human activities on the systems of Planet Earth.
Chapter 17 of Silent Spring, the magnificent book of Rachel Carson (1962), begins with
a vigorous presentation of our responsibilities: “We stand now where two roads diverge. But
unlike the roads in Robert Frost’s familiar poem, they are not equally fair. The road we have
long been travelling is deceptively easy, a smooth superhighway on which we progress with
great speed, but at its end lays disaster. The other fork of the road – the one less travelled
by – offers our last. Our only chance to reach a destination that assures the preservation of our
Earth. The choice, after all, is ours to make”.

6 WATER SIGNALS

Rachel Carson advice has been followed and new water policies have been implemented
all over the world. However, the combined demands of agriculture, new urban settlements,
an expanding industry and a growing population widen the gap between preservation and
disturbance of natural waters.
Several processes affecting water resources are far from equilibrium and may serve as
signals of water mismanagement, pointing out to fresh technical undertakings and coming
political issues. To mention but a few, see Table 1.
Human impact on the biosphere waters was responsible for the expansion of many biotopes
and ecosystems and the creation of bridges among them. Table 2 identifies a few examples.
Table 1 of water signals focuses on negative impacts. To balance the view, the social benefits
of water exploitation must be incorporated, waving a new list of water signals.
Living standards, health, labour conditions, life expectancy, all improved with the era of
economic development started by Industrial Revolution. The appreciation for springs, cas-
cades, rivers, lakes and the maritime littoral has spread to most of the contemporary world
population. Water resorts and water sports have became commonplace in Europe where lit-
toral cities witnessed a steady growth during the second half of the 20th century. However,
many human communities are lagging behind others in economic and social development and
human welfare.
Water and nature. The berth of life 249

Table 1. Signals of water mismanagement due to several processes affecting water resources.

Water signals Comments

Eutrophication of continental waters Diffuse contamination from agriculture is


growing as chemical fertilizers are more
widely used.
Leachates from uncontrolled rubbish dumps
find their way into waters.
Waste water treatment plants rarely controls
P in effluents sufficiently to prevent water
enrichment. 1–2 mg/L concentrations in
effluents or an 80% reduction in wastewater
have been set as an European standard for
some countries.
Eutrophication of littoral waters As population shifts to the sea shore, point
sources of untreated waste waters released
into the sea keep growing.
Tourist settlements with fluctuating
occupation rarely possess waste water
treatment plants adequate to the
variable load.
In poor regions the water works design
usually includes sewage collection and its
deposition in rivers or coastal waters
with no treatment.
New organic molecules are added to urban Waste water treatment cannot prevent the
waste water in significant amounts such as growing concentration of biologically active
hormones, beta-blockers, antibiotics, or molecules in effluents.
tensioactives. Hormones at concentrations found in
wastewater have measurable effects on
aquatic organisms such as toads, inducing
significant development alterations.
Growing water demands from New demands tend to be met by more
agriculture, industry and urban abstraction from aquifers and increased
developments. Water requirements river or lake pumping. River impoundment
may be close to 10,000 km3 /yr, for watershed regulation starts far reaching
roughly 1/10 of continental changes in river ecology.
precipitation. In large rivers with important sediment loads,
The increase of rent in population and the the loss of sedimentation unbalances coastal
development of urban water supplies areas close to the estuary, also
greatly increase water demands in an lowering biological productivity in them.
early development phase. Overexploitation of aquifers has several
consequences other than consumption of water
resources. Desiccation of connected springs,
upwellings and wetlands. Loss or reduction of
habitats and populations, generally reducing
biodiversity.
In littoral areas, aquifer salination occurs
due to marine water intrusion.
Repeated introduction of alien fish In some cases, such as the Nile perch (Lates
species in continental waters for sport niloticus) in Victoria Lake, led to a biodiversity
or commercial fisheries. collapse. In many other cases, native fish
(Continued )
250 F. García Novo & F. García Bouzas

Table 1. (Continued)

Water signals Comments

populations retreat when aggressive newcomers


compete for resources or prey on them.
Intense trade of aquatic organisms with Other than introductions, there is a problem
little control. Domestic aquaria house with illegal trade of protected species and
many menacing invaders. the destruction of populations of tropical native
species of fishes and amphibians.
The use of rivers and lakes as cooling A rise in temperature favours the introduction
circuits for power plants creates long- of tropical or subtropical organisms into
standing heated plumes or raised river cool or temperate regions.
temperatures. Nuclear plants raise water The addition of chemicals to cooling circuits
concentration of tritium and other isotopes, to prevent animal growth damages biota on the
some of them causing health impacts. marine littoral and in lakes and rivers.
Untreated sludge from chemical plants Ecological concentration mechanisms can raise
may introduce to continental waters harmless concentrations of chemicals in the
some elements, ions or molecules rarely environment to significant or toxic levels
confronted by ecosystems, causing along trophic chains. Hg concentration in
harmful ecological effects. Minamata Bay (Japan) from water to fish and
Desalination plants release a concen- sea food in human diet was an early warning
trated brine creating a plume of hyper of the process.
saline waters close to the seashore. Large discharges of waste water, even when
contamination was caused by organic matter or
naturally abundant compounds (such as nitrates,
nitrites, ammonia, phosphates), easily exceeds
regulatory ability of aquatic systems, causing
alterations and the eventual collapse of the
ecosystem.
Mineral oil can be dispersed, oxidised,
degraded, and assimilated by natural systems, in
moderate quantities. But point concentrations
may be very high in oil fields, chemical
plants, or after accidental releases from tankers
and pipelines. It is oil concentration what
makes ecological disasters, fouling soils, rivers
and coastal areas. Hyper saline waters are
harmful to most marine species.
World biotopes are sensitive to interfer- Biotopes, populations, aquatic ecosystems,
ence in the water cycle: overexploitation natural landscapes and cultural landscapes
of aquifers feeding wetlands, reduction of response to interventions are most varied from
river flow and sediment transport, fragmen- small readjustments to collapse and substitution
tation of rivers with dams and impound- by a different system.
ments, drainage of wetlands, eutrophication The release of species from other
and contamination of continental waters. biogeographical regions often aggravates the
And above all, the climatic change. situation of local endemic taxa.
Human cultures have created a rich legacy New balances among cultural survival, health,
on water issues. Technical changes in the comfort and efficient use of resources must be
exploitation and loss of traditional developed.
resources (such as fish, flax, watermills), The paradigm of sustainability needs profound
severs cultures at a world scale. reshaping of traditional cultures and, the more
so, of contemporary western technical culture.
Sustainability of water resources merits an
independent assessment.
Water and nature. The berth of life 251

Table 2. Ecological changes induced by water resource exploitation.

Process Comments

Growing Hot deserts, range and dry lands show limited production and biomass due to
water water scarcity. Irrigation favours crop and husbandry productivity as well as
demands the expansion of associated wild species of forbs and soil invertebrates.
When large irrigation schemes are applied to hot countries there is a drop in
peak temperatures in the area. Sometimes irrigation helps introduce more
advanced agriculture practices and protects soils from desertification.

River Dams interrupt the river continuum and insert lakes (reservoirs) in it. Still
impoundment stratified waters with reduced suspended solid load and deep columns (as
compared to rivers) offer new habitats for aquatic species. Birds, insects,
reptiles and amphibian faunas of river and reservoirs differ. Fish communities
of lentic and lotic systems are markedly different.

The release The shifting of aquatic species across terrestrial or oceanic barriers,
of species insurmountable to them as individuals, opened up a new era to biodiversity.
from other Fish species introduced to adequate media may expand adding to the local
biogeographical diversity. Introductions are deleterious to some native wildlife, but often
areas favours favour other species until a new equilibrium is reached.
introductions In the future, world continental waters will probably lose most local wildlife
species but will be enriched by fast growing species coming from every
continent.
Some introductions bring important benefits to local inhabitants: edible fish
for commercial fishing and sport. Fast growing fish and sea food for
aquaculture. Filtering fish to abate water eutrophication in lakes. Mosquito
fish Gambusia to fight malaria. Others introductions are detrimental to
humans such as parasites (Bilharzia) or aquatic pests.

Urban Modern cities have created new aquatic environments that have been
expansion colonised by wild species.
offers new Pipes carrying irrigation or river waters may exhibit a thick cover with an
habitats to association of filtering animals such as Hydroidea (Cordilophora
many wild caspia) and molluscs (Mytilopsis leucophaeta, Corbicula fluminea).
species Sewers offer a network of channels and pipes with a steady flow of water rich
in organic matter and with limited temperature fluctuations. Decomposers
thrive in this warm, moist and well fed environment, such as microbes,
protozoa, fungi and some worms. And larger animals such as insects and rats.
Tropical cockroaches have adopted sewers as an expansion of their habitat all
over the world. Small alligators or crocodiles, aquatic snakes, otters, some
fish as mullet (Liza) species, survive in large sewers. Sources of disturbance
are the occasional water flashes caused by an intense precipitation, and the
release of chemicals to sewers.

Air Air filters collect particles and condensation provides water in devices with
conditioning large surfaces, offering an aerated water medium enriched in organic matter.
and new This environment suits well Legionellacea, a large family of bacteria often
urban causing epidemic outbreaks of respiratory diseases. This happens when
environments microbes are dispersed to the air in droplets from the air conditioning systems.
Water pockets in roofs, collectors, pipes, offer a temporary water environment
that can be exploited by bacteria and algae. Basements, wells, gardens,
fountains, river banks, piers and harbours, boats, offer an expanded array of
wet urban habitats to biological forms.

(Continued )
252 F. García Novo & F. García Bouzas

Table 2. (Continued)

Process Comments

Navigation Boat hulls behave like wandering rocks for marine sessile organisms. On some
occasions, the distribution of marine species from shallow waters over deep
waters was made possible through this mechanism. Caulerpa taxifolia has been
expanding on the Mediterranean from harbour to harbour likely fixed on sport
sailing boats. Enteromorpha intestinalis, a marine green algae, has reached
Switzerland and colonised the Rhine River with freshwater forms fixed to
commercial boats operating in this water way.
Mississippi River is behaving as a continental loop for freshwater species from
the Great Lakes to the Caribbean Sea.
When ships in ballast emptied tanks in harbours the waters often contain live
specimens from distant areas that can be introduced. There is growing concern
about the aggressive Decapoda (crustacean) species introduced in fluvial ports
and marine areas through this mechanism.
Used tyres and metal scrap cargo in ships is known to offer water pockets for
reproduction of mosquito species and their introduction to new areas.

We are forcing the water cycle to meet human needs often overlooking the basic prin-
ciples governing the life of organisms, the assembly of their communities, the organization
of ecosystems and the multiple roles of water in the biosphere. The onset of climatic change
dramatically signals how far human interference pervaded the water cycle of biosphere; and
the risk of severing the unity of natural waters.
Sustainability policies must take full advantage of self regulatory processes in natural
systems, bringing them to social uses by means of new technologies founded on hydrology,
limnology and ecology. As Ian McHarg advised, long ago (1969): Design with Nature!

REFERENCES

Belnap, J. & Lange, O.L. (2001). Biological soil crusts: Structure, function and management. Springer–
Verlag, Berlin. 503 pp.
Carson, R. (1962). Silent Spring. Houghton Mifflin. Boston, USA.
Cloudsley-Thompson, J.L. (1991). Ecophysiology of Desert Arthropods and Reptiles. Springer–Verlag,
Berlin. 216 pp.
Constanza, R.; d’Arge, R.; de Groot, R.; Farber, S.; Grasso, M.; Hannon, B.; Limburg, K.; Naeem, S.;
O’Neill, R.V.; Paruelo, J.; Raskin, R.G.; Sutton, P. & van den Belt, M. (1997). The value of the world’s
ecosystem services and natural capital. Nature, 387: 253–260 (May 15).
Erwin, T. (1983). Beetles and other insects of tropical forest canopies at Manaus sampled by insecticidal
fogging. In: S.L. Sutton; T.C. Whitmore & A.C. Chadwick (eds.). Tropical rain forest: ecology and
management. Blackwell, London, UK: 59–75.
Herrera, C.M. (1992). Historical effects and sorting processes as explanation for contemporary eco-
logical patterns: character syndromes in Mediterranean woody plants. The American Naturalist, 140:
421–446.
McHarg, I. (1969). Design with Nature. Doubleday and Company. New York, USA.
Mojzsis, S.J.; Arrhenius, G.; McKeegan, K.D.; Harrison, T.M.; Nutman, A.P. & Friend, C.R.L. (1996).
Evidence for life on Earth before 3800 million years ago. Nature, 384: 55–59 (November 7).
Wittfogel, K. (1957). Oriental despotism: a comparative study of total power. Yale University Press.
New Haven, Connecticut, USA.
CHAPTER 15

Water and nature: a critical link for solving the water


management crisis

G. Bergkamp
IUCN – The World Conservation Union, Gland, Switzerland

ABSTRACT: Environmental aspects of water management have long been neglected. During the last
two decades, however, progress has been made in specific cases. More recently, a range of practical tools
and mechanism have come to be used to incorporate the values of ecosystem services in water resources
planning, decision-making and management. These include environmental flow management, economic
valuation of ecosystem goods and services and payments for maintenance of environmental services.
Nature, by using these new approaches, is becoming part of solving the water management crisis.

Keywords: Water and nature, environmental flows, ecosystem valuation, ecosystem services,
protection of wetlands

1 INTRODUCTION

Early whistle-blowers of the degradation of the world’s water resources, like Aldo Leopold and
Rachel Carson, are emblematic for the way the modern world has dealt with its relationship
with water and nature. Since the publication of Carson’s book (1962), thousands of dams have
been constructed worldwide. Today, 70% of rivers are fragmented by large scale infrastructure,
affecting migration of fish and other species (Bergkamp et al., 2000). During the last 100
years more than 50% of the world’s wetlands have been drained and turned into agricultural
production areas. Up until the 1980s most developed and developing societies alike choose to
ignore the early warning signs related to a mismanagement of the world’s water resources. They
preferred to develop their water resources in ways that would maximise short-term economic
growth rather than safeguard their resource base for future generations.
During the last two and a half decades, societies have started to respond, however, to
signals of degradation. Especially in the developed world, more and more information became
available on the seriousness of the situation. This triggered policy-level responses. For example,
it was in the late 1980s and early 1990s that with the Brundtland report, Caring for the Earth
and Agenda 21, environmental issues and sustainable development were put on the policy
agendas at the national level. At that time, the water society at large was still mainly focused
on a Green Revolution (1960–1970) and the Water Supply Decade (1981–1990). Only in the
early 1990s, with the Dublin Principles and Chapter 18 of Agenda 21, a start was made
catalysing change towards a more sustainable management of water resources.
Since the early 1980s, significant changes have occurred in the environmental aspects of
water management. With increased knowledge about the effects of the development of water
254 G. Bergkamp

resources and industries came a growing demand for reducing those impacts. Some significant
improvements in water quality and in the protection of wetlands were made. For example, in
OECD countries, pollution of rivers and lakes with phosphate has been reduced by 40%. Since
the early 1970s, more than hundreds of thousands of hectares of wetlands have been protected
under the Ramsar Convention. These changes are steps in the right direction.
Much of these improvements were built upon an improved understanding of causes of
pollution and degradation, and political will. The latter was often driven by motives that were
not necessary environmental in the first place. For example, the deteriorated quality of bathing
water in many places in Western Europe caused great public concern, which caused politicians
to act. Water quality standards and monitoring were put in place, while new water treatment
systems were built and existing ones upgraded.

2 ENVIRONMENTAL FLOWS: A NEGOTIATED APPROACH TO RIVER FLOW


MANAGEMENT

Water resource managers are now increasingly coming to terms with the need to take a holistic
view of the river ecosystem. This includes taking care of aquatic ecosystems and the resources
they provide for long-term economic viability and allocating appropriate amounts of water
to them.
Ecosystems provide goods and services to humankind which are relevant to water supplies.
Upper watershed forests regulate water flow, ensure recharge of aquifers and reduction of silt
loads in rivers. Wetlands can purify water and filter out chemical substances harmful to humans
or crops. Rivers and lakes also provide sources of protein through fish and other animals. Often
these services are of direct benefit to local communities living from the resource (IUCN, 2000).
The concept often used for this allocation is environmental flows. It is defined as the water
regime provided within a river, wetland or coastal zone to maintain ecosystems and their
benefits, where there are competing water uses and where flows are regulated. Environmental
flows must be seen within the context of applying Integrated Water Resources Management
(IWRM) in catchments and river basins. Environmental flows will only ensure a healthy river,
however, if they are part of a broader package of measures, such as soil protection, pollution
prevention, and protection and restoration of habitats (Dyson et al., 2003).
Taking steps to manage for environmental flows brings into focus the struggle over access to
and ownership of water and water rights. In systems where water is already over-allocated, the
challenge of environmental flows may include reallocating or conserving water from existing
private users and returning it to the river. Before starting to work on environmental flows, one
therefore needs to realise that a wide range of stakeholders will have to be involved.
The degree of good health at which the river will be sustained is a societal judgement that
will vary from country to country and region to region. What the appropriate environmental
flow is for a particular river will thus depend on the values for which the river system is to
be managed. Those values will determine the decisions about how to balance environmental,
economic and social aspirations and the uses of the river’s waters.
To set an environmental flow, one needs to identify clear stream objectives as well as water
abstraction and use scenarios. Objectives need to have measurable indicators that can form the
basis for water allocations. Useful objectives can be, for example, maintaining the brown trout
at 1995 levels, preserving at least 75% of downstream mangrove forests, or maintaining river
Water and nature: a critical link for solving the water management crisis 255

nitrate levels below a particular standard. Where a system is seriously over-committed and
values do not allow a sufficient reallocation of resources to restore the entire system, certain
river stretches or wetland sites may be targeted for protection and specific water allocations.
For rivers with high biodiversity values, for example, an environmental flow might be provided
to preserve the natural state of the river system.
An environmental flow programme will not necessarily have ecological gains as the only or
even the primary outcome. It will need to strike a balance between water allocations to satisfy
the ecological water requirement and other water use needs like those of hydropower gener-
ation, irrigation, drinking water or recreation. Developing an environmental flow programme
therefore means articulating the core values on which to base decisions, determining what
outcomes are sought and defining what trade-offs those will entail. In this way, environmental
flow setting and implementation provide an opportunity for basin and watershed stakeholders
to start discussing the existing allocation patterns that might not be sustainable.

3 COUNTING ECOSYSTEMS AS PART OF WATER SERVICE INFRASTRUCTURE

Ecosystems provide a wide range of economic goods and services for human production and
consumption, such as fish, timber, fuel wood, medicines and pasture. As pointed out above,
water allocations need to be made to maintain those services. On the supply side, ecosystems
such as forests and wetlands generate economic services which maintain the quality and
quantity of water available for human use or mitigate flooding and drought. Healthy ecosystems
are thus important to secure water supplies, while water supplies to maintain ecosystem health
are critical.
A recognition of the relationship between ecosystems status and water service infras-
tructure has long been missing from water resources planning and infrastructure investment
programmes. Economic arguments are, however, an important determinant of how water is
allocated and used and how priorities for water investments are set. To date little appreciation
exists for the fact that ecosystems are economic users of water and form equally a critical part of
the water supply chain. Though they are essential for water service delivery, they often remain
a sharply under-funded part of investment in the water sector (Emerton & Bos, in prep.).
To improve on this practice, it is critical to determine the benefits from ecosystems in mon-
etary terms. Economic valuation of ecosystem benefits can be a powerful tool to incorporate
those values into economic planning and decision-making. It provides a basic and compar-
able indicator of economic values which can be used to judge ecosystem use alongside other
economic sectors and activities.
A wide range of techniques now exist to value different components of the total economic
value of ecosystems. The total economic value includes the direct values, e.g. of products such
as fish and timber, and the indirect values, e.g. of services such as flood control and nutrient
retention. They also include option values, the premium placed on maintaining resources for
future use. Finally they include the intrinsic values of resources and ecosystems irrespective
of their use, such as cultural and spiritual significance.
Available techniques for determining ecosystem values include, for example, determining
the market-price of ecosystems goods and services, and valuing the effect of change on
production output of an ecosystem. A range of experiences is now available around the world.
For example, in Guatemala the effect of rainforest degradation on stream flow and agricultural
256 G. Bergkamp

profits was found to be worth US$ 15,000–53,000 in two river basins (Brown et al., 1996). In
the Zambezi river basin, estimating the total value of its major wetland areas indicated that they
have a marginal value of US$ 145 million/yr, or an average of US$ 48/ha (Seyam et al., 2001).
A major challenge arises once values have been determined. How to ensure those values are
used in decision making? How to make use of those values when decision need to be made to
developing a hydropower scheme and preserving river values for downstream users? Several
techniques exist to translate these data into information, indicators and criteria that can be used
to weigh different development options. They include, for example, cost-benefit analysis and
multi-criteria analysis. In Northern Cameroon these techniques were used to determine the
cost and benefits of restoration of the 3000 km2 Waza Logone floodplain. The study showed
that the cost of not having an annual floodplain inundation is more than US$ 2 million/yr
due to loss of fish, fodder and grazing grounds. Restoring the floods could be done at an
investment cost of US$ 10 million/yr, which would result in benefits of US$ 2.5 million/yr to
the region’s economy. The investment would pay itself back in five years (IUCN, 2001).
Valuing the benefits and the costs of ecosystems goods and services can thus provide a
meaningful contribution to solving the water management crisis. By determining these values
and incorporating them into planning and decision-making, more optimal water investment
programmes can be designed and implemented.

4 PAYMENT FOR ECOSYSTEM SERVICES

Increasingly, potential beneficiaries of ecosystem services make payments to ecosystem man-


agers. These payments reward land owners and resource users for managing land and water
resources in a way most favourable to the protection of the ecosystem services of interest. The
majority of compensation mechanisms in place have been established by the public sector,
which has been buying services such as maintenance of hydrological flow from private land
users. Many of the existing schemes relate ecosystems services directly to the provision of
water supplies to cities, hydropower stations or irrigated farmland.
A well known example is the NewYork City Water Supply that started to invest in watershed
protection as a cost-effective means to comply with water quality regulations. The City receives
its water supply from two watersheds located 125 miles (200 km) north of the City. Over a ten
year period, the city invested US$ 1500 million to protect the watershed natural purification
processes. Measures included US$ 472 million to control point-source pollution through the
upgrade and rehabilitation of city-owned sewage treatment plants, US$ 300 million for land
acquisition, and US$ 40 million for Best Management Practices in agriculture and forestry. The
annual costs of operating the watershed protection program have been reduced to US$ 20–30
million, well below the operational costs of a filtration facility of US$ 6000–8000 million plus
an estimated US$ 200–300 million in annual operating expenses (Perrot Maître & Davis, 2001).
In many parts of the world, hydropower companies have decided to make voluntary con-
tributions to watershed protection. This is the case for example in Quito, Ecuador, where
since 2000, the water utility and the hydroelectric utility have decided to transfer 1% of their
revenues from water and electricity sales into a Trust Fund that will be used for watershed
protection. It is expected that over US$ 2 million/yr will be generated.
In Costa Rica, five hydroelectric companies are investing in watershed protection. They
annually allocate US$ 10–42/ha or over US$ 600,000/yr to a government fund which is used
Water and nature: a critical link for solving the water management crisis 257

to compensate forest owners for forest protection activities. Companies do so for two reasons:
(1) they expect, through watershed protection, to ensure a regular supply of water throughout
the year; and (2) they can increase revenues from electricity production and reduce dredging
costs associated with reservoir sedimentation.
Bottling companies are also heavily dependent on a secure water supply of good quality.
Protecting the resources is a fundamental element of their success. Over the last years, a number
of companies have bought land around their springs to reduce the risks of water contamination
by outside activities. For example, Nestlé (Vittel S.A.) invested US$ 9 million to purchase land
around the Vittel springs and provided a US$ 24.5 million of subsidies to land users to adopt
Best Management Practices. In doing so, the company is investing in protecting the water
purification services provided by forest and pasture lands. Research at the French National
Agronomic Institute (INRA) demonstrated that the programme is cost-efficient as long as one
hectare of well-managed pasture is able to filter and produce 3000 m3 /yr of drinkable water.

5 CONCLUSIONS

The above examples show practical approaches that make ecosystems services part of the
solution, rather than being perceived as part of the problem. If we are to finance the sustainable
management of our water resources, we will need to convincingly demonstrate that investing
in their protection and wise use makes good economic and financial sense. Compensating
those land and water managers that protect the resource and maintain a critical element of the
water supply chain is likely to become more common practice in the future.

REFERENCES

Bergkamp, G.; McCartney, M.; Dugan, P.; McNeely, J. & Acreman, M. (2000). Dams, Ecosystem Func-
tions and environmental restoration. World Commission on Dams Thematic Reviews – Environmental
Issues II.1. World Commission on Dams, Cape Town, South Africa: 1–186.
Brown, M.; de la Roca, I.; Vallejo, A.; Ford, G.; Casey, J.; Aguilar, B. & Haacker, R. (1996). A Valuation
Analysis of the Role of Cloud Forests in Watershed Protection: Sierra de las Minas Biosphere Reserve,
Guatemala and Cusuco National Park, Honduras. RARE Center for Tropical Conservation, Fundación
Defensores de la Naturaleza and Fundación Ecológica.
Carson, R. (1962). Silent spring. Houghton Mifflin, New York, USA.
Dyson, M.; Bergkamp, G. & Scanlon, J. (eds.) (2003). Flow – The essentials of environmental flows.
IUCN, Gland, Switzerland: 1–118.
Emerton, L. & Bos, E. (in prep.). Value – Counting ecosystems as an economic part of water
infrastructure. IUCN, Gland, Switzerland.
IUCN (2000). Vision for Water and Nature. A world strategy for Conservation and sustainable manage-
ment of water resources in the 21st century. World Water Vision – Environment component. World
Water Council/IUCN. Gland, Switzerland: 1–52.
IUCN (2001). Economic Value of Reinundation of the Waza Logone Floodplain, Cameroon. Projet de
Conservation et de Développement de la Région de Waza-Logone, Maroua, Cameroon.
Perrot-Maître, D. & Davis, P. (2001). Case studies: Developing Markets for Water Services from Forests.
Forest Trends, Washington, D.C., USA. Web Page: (http://www.forests-trends.org).
Seyam, I.M.; Hoekstra, A.Y.; Ngabirano, G.S. & Savenije, H.H.G. (2001). The Value of Freshwater
Wetlands in the Zambezi Basin. Paper presented at Conference on Globalization and Water Resources
Management: the Changing Value of Water. AWRA/IWLRI–University of Dundee.
VIII
New Technologies to Cope with
Water Scarcity
CHAPTER 16

Water recycling – A relevant solution?

T. Asano
Department of Civil and Environmental Engineering, University of California, Davis, USA

ABSTRACT: Water reclamation, recycling, and reuse have become important elements in integrated
water resources management. The role of water recycling and reuse is reviewed in this chapter with
respect to the evolution of the practice, applications, technology, health and regulatory requirements,
and planning. The future of water recycling and reuse are also discussed in the context of a relevant
solution to water shortages and water pollution.

Keywords: Planning, public health, recycling, water resources management, water reclamation,
wastewater treatment, water reuse

1 INTRODUCTION

For water supplies to be sustainable, the rate at which water is withdrawn from water sources
needs to be in balance with the rate of renewal or replenishment of these water sources. In addi-
tion to a balance of water quantity, water quality must also be sustainable or recoverable. Precip-
itation serves to replenish water sources by augmenting surface water supplies and recharging
groundwater. Shifts in land use patterns (e.g. urbanization, agriculture, dams and reservoirs)
alter the rate, extent, and spatial distribution of freshwater withdrawal, consumption, and
replenishment. Water that is withdrawn for societal needs is also a potential source of water
replenishment that should be considered in the sustainability equation (Levine & Asano, 2004).
The sustainability of water resources is of particular importance in light of projected
increases in global population. It has been reported that the current world population of 6200
million is increasing at a rate of about 1.2% per year (United Nations, 2003) with the highest
rates of population growth occurring in urban areas where supplies of freshwater tend to be
limited. On a global scale, about 3800 km3 /yr of water is withdrawn to meet societal needs.
Increasing urbanization has resulted in an uneven distribution of population and water, thus
imposing unprecedented pressures on limited water supplies. These pressures are exacerbated
during periods of drought. When viewed on a global scale, the total available volume of renew-
able freshwater is several times more than is needed to sustain the current world population.
However, only about 31% of the renewable water is directly accessible to population centers
due to geographical constraints and seasonal variations (Postel, 2000).
Historically, after water has been used for societal needs, it has been labeled as waste
water and treated to the extent perceived necessary for discharge into receiving water or on
land. During most of the 20th century, the emphasis of wastewater treatment was on pollution
abatement, protection of public health, and prevention of environmental degradation through
removal of biodegradable material, nutrients, and pathogens. However, over the last few
262 T. Asano

decades, the potential for recovering water from [waste]water has been recognized. In fact, in
many parts of the world, it is no longer practical or possible for water to be used only once.
Reclaimed water is a valuable water resource for non-potable and subpotable water uses. By
using treated wastewater effluent to augment existing water supplies, water resources can be
conserved more economically and effectively. Water recycling and reuse, in conjunction with
conservation, efficient management and use of existing water resources results in a sound and
integrated approach to water resources management. In many industrialized nations there are
growing problems associated with developing adequate water supplies in an environmentally
sound manner. Water shortages, especially during periods of drought, have made it necessary
for many municipalities to mandate the reduction of water consumption. In addition, the costs
of municipal and industrial wastewater disposal for water quality protection and pollution
abatement are increasing. In developing countries, particularly those in arid parts of the world,
reliable low cost, low technology methods are needed for acquiring new water supplies and
protecting existing water sources from pollution.
In the planning and implementation of water reclamation and reuse, the intended water reuse
applications dictate the extent of wastewater treatment required, the quality of the finished
water, and the method of distribution and application. As technology continues to advance
and the reliability of water reuse systems is demonstrated, it is likely that water recycling and
reuse will continue to expand in the future and become a more relevant solution in the context
of integrated water resources management.
The purpose of this chapter is to provide an introduction to the major elements of water
reclamation, recycling, and reuse with an emphasis on significant developments that have
paved the way for present and future applications. The major topics covered are: (1) evolution
of water reuse; (2) water recycling and reuse terminology; (3) the role of water recycling
in the water cycle; (4) water reclamation and reuse benefits; (5) water reuse applications;
(6) overview of wastewater treatment technology; (7) health and regulatory requirements;
(8) water reuse planning; and (9) unresolved problems and research needs. The goal of this
chapter is to stimulate the readers a new thinking in water resources management and to
illustrate the relevant role played by water reclamation, recycling, and reuse in water resources
planning and management.

2 EVOLUTION OF WATER REUSE

Indications for utilization of wastewater for agricultural irrigation extend back approximately
5000 years in Crete, Greece (Angelakis & Spyridakis, 1995). In more recent history, during
the 19th century, the introduction of large-scale wastewater carriage systems for discharge
into surface waters led to indirect use of sewage and other effluents, downstream, for inad-
vertent potable water supplies. This unplanned reuse, coupled with the lack of adequate water
and wastewater treatment, resulted in catastrophic epidemics of waterborne diseases such as
Asiatic cholera and typhoid during the 1840s and 1850s. However, when the water supply link
with these diseases became clear, engineering solutions were implemented that included the
development of alternative water sources using reservoirs and aqueduct systems, the relocating
of water intakes upstream and wastewater discharges downstream as in the case of London,
and the progressive introduction of water filtration during the 1850s and 1860s (Young, 1985;
Barty-King, 1992).
Water recycling – A relevant solution? 263

2.1 Water reuse in the USA and the European Union


The development of programs for planned reuse of wastewater within the USA began in the
early part of the 20th century. The State of California pioneered efforts to promote water
reclamation and reuse. The city of Bakersfield has used reclaimed water since 1912 to irrigate
corn, barley, alfalfa, cotton, and pasture. The first reuse regulations were promulgated in
1918 by the State of California. Some of the earliest water reuse systems were developed to
provide water for irrigation with projects implemented in both Arizona and California in the
late 1920s. In the 1940s chlorinated wastewater effluent was used for steel processing, and in
the 1960s urban water reuse systems were developed in Colorado and Florida.
During the last quarter of the 20th century, the benefits of promoting water reuse as a means
of supplementing water resources have been recognized by most state legislatures in the USA
as well as by the European Union. For example, in 1970 the California State Water Code stated
that “it is the intention of the Legislature that the State undertake all possible steps to encourage
development of water reclamation facilities so that reclaimed water be available to help meet
the growing water requirements of the State” (California Water Code, 1988 Amendments).
In the same context, the European Communities Commission Directive (91/271/EEC) declared
that “treated wastewater shall be reused whenever appropriate. Disposal routes shall minimize
the adverse effects on the environment” (EEC, 1991).
In recent years, increased interest in water reuse in many parts of the world is occurring
in response to growing pressures for high quality, dependable water supplies by agriculture,
industry and the public; a situation that is exacerbated by rapid population growth in urban
areas, drought and other climatic irregularities. Today, technically proven wastewater treatment
and water purification processes exist to produce water of almost any quality desired.

2.2 Milestone events


Milestone events that have been significant for the evolution of water reclamation, recycling,
and reuse are itemized on the timeline in Figure 1. Microbiological advances in the late
19th century precipitated the Great Sanitary Awakening (Fair & Geyer, 1954) and the advent
of disinfection processes. The development of the activated sludge process in 1904 was a
significant step towards advancement of wastewater treatment and pollution control and the
development of biological treatment systems. In 1918, the California State Board of Public
Health adopted its first regulations addressing the use of sewage for irrigation.
Technological advances in physical, chemical, and biological processing of water and
wastewater during the early part of the 20th century led to the Era of Wastewater Reclama-
tion, Recycling, and Reuse. Since the 1960s, intensive research efforts, fueled by regulatory
pressures and water shortages, have provided valuable insight into health risks and treatment
system design concepts for water reuse. In 1965, the Israeli Ministry of Health issued reg-
ulations to allow the reuse of secondary effluents for crop irrigation with the exclusion of
vegetable crops that are eaten uncooked. In 1968 extensive research on direct potable reuse
was conducted in Windhoek, Namibia. In the USA, a milestone event was the passage of the
Federal Water Pollution Control Act in 1972 (later renamed: the Clean Water Act) “to restore
and maintain the chemical, physical, and biological integrity of the Nation’s waters” with the
ultimate goal of zero discharge of pollutants into navigable, fishable, and/or swimmable waters.
During the 1970s and 1980s several comprehensive research and demonstration projects were
conducted to evaluate the potential for non-potable and potable reuse with a major emphasis
264 T. Asano

EARLY WATER AND SANITATION SYSTEMS: 3000 BC to 1850

Minoan
Civilization
• 97 AD--Water Supply Commissioner for City of Rome-Sextus Julius Frontius
Sewage farms in Germany
Sewage farms in UK
Legal use of sewers for human waste disposal:
London (1815), Boston (1833), Paris (1880)

Cholera epidemic in London


(also 1848-49 and 1854)
Sanitary status of Great Britain Labor Force: Chadwick Report
"The rain to the river and the sewage to the soil"

3000 BC 1550 1600 1650 1700 1750 1800 1850

GREAT SANITARY AWAKENING: 1850 to 1950


Cholera epidemic linked to water pollution control by Snow (London)
Typhoid fever prevention theory developed by Budd (UK)
Anthrax connection to bacterial etiology demonstrated by Koch (Germany)
Microbial pollution of water demonstrated by Pasteur (France)
Sodium hypochlorite disinfection in UK by Down to render the water "pure and wholesome"
Chlorination of Jersey City, NJ water supply (USA)
Disinfection kinetics elucidated by Chick (USA)
Activated sludge process demonstrated by Ardern and Lockett in UK
First regulations for use of sewage for irrigation purposes in California

1850 1870 1890 1910 1930 1950

ERA OF WASTEWATER RECLAMATION, RECYCLING AND REUSE: POST 1960


California legislation encourages wastewater reclamation, recycling and reuse
Use of secondary effluent for crop irrigation in Israel
Research on direct potable reuse in Windhoek, Namibia
US Clean Water Act to restore and maintain water quality
Pomona Virus Study; Pomona, CA
California Wastewater Reclamation Criteria (Title 22)
Health effects study by LA County Sanitation Districts, CA
Monterey Wastewater Reclamation Study for Agriculture, CA
WHO Guidelines for Agricultural and Aquacultural Reuse
Total Resource Recovery Health Effects Study;
City of San Diego, CA

Potable Water Reuse Demonstration Plant; Denver, CO


Final Report -- plant operation began in 1984

1960 1965 1970 1975 1980 1985 1990 1995 2000

Figure 1. Milestone events in the evolution of wastewater treatment, reclamation, recycling, and reuse
(Adapted from Asano & Levine, 1996, 1998).

on quantifying health risks and defining treatment and technological requirements. These
research efforts have resulted in increased implementation of water reuse projects in various
regions and the evolution of new reuse alternatives.
The continued testing and implementation of treatment systems and new applications have
helped to overcome many technical barriers to water reuse projects. Improvements in treatment
process reliability, risk assessment, and public confidence in reuse systems in conjunction with
Water recycling – A relevant solution? 265

increasing water demands and pollution control requirements have promoted the integration
of water reuse into water resources management strategies throughout the world.

3 WATER RECYCLING AND REUSE TERMINOLOGY

The terminology used in water reclamation, recycling, and reuse has evolved from sanitary
and environmental engineering practice. The water potentially available for reuse includes
discharge from municipalities (sewage, municipal wastewater), industries (industrial wastes),
agricultural return flows, and storm water. Of these, return flows from agricultural irrigation
and storm water are usually collected and reused without further treatment.
Water reclamation involves the treatment or processing of wastewater to make it reusable,
and wastewater reuse or water reuse is the beneficial use of the treated water. In recent years,
term water recycling and reuse, instead of wastewater, is preferred from the public acceptance
point of view. Reclamation and reuse of water frequently require water conveyance facilities
for delivering the reclaimed water and may require intermittent storage of the reclaimed water
prior to reuse.
Reclaimed water can be used directly (direct reuse) without passing through a natural body
of water for applications such as agricultural and landscape irrigation. Indirect reuse includes
mixing, dilution, and dispersion of reclaimed water by discharge into an impoundment, receiv-
ing water, or aquifer prior to reuse. Indirect reuse, through discharge of a treated effluent to
receiving water for assimilation and withdrawal downstream, while important, does not nor-
mally constitute planned reuse. For example, the diversion of water from a river downstream
of a discharge of treated water constitutes an incidental or unplanned reuse. The topics covered
in this chapter are related to deliberate or planned reuse as defined above.

3.1 Unplanned indirect water reuse


Unplanned indirect water reuse, through effluent disposal to streams and groundwater basins,
has been a long accepted practice throughout the world. Many communities situated at the
end of major waterways: New Orleans, USA (the Mississippi River); London, England (the
River Thames); the cities and towns in the Rhine River Valley, Germany; and Osaka, Japan
(the Yodo River); ingest water that has already been used many times through repeated river
withdrawal and discharge. For example, it has been estimated that more than 80% of the
water in the Santa Ana River in Southern California originates from wastewater discharged by
upstream municipal wastewater treatment plants. Similarly, river beds or percolation ponds
may recharge underlying groundwater aquifers with waste-containing water, which in turn is
withdrawn by subsequent communities for domestic water supply.
It is also increasingly important to differentiate between potable and non-potable reuse
applications. Potable water reuse refers to the use of highly treated reclaimed water to augment
drinking water supplies. Direct potable reuse is the incorporation of reclaimed water into a
potable water supply system, without relinquishing control over the resource. Indirect potable
reuse includes instead an intermediate step in which reclaimed water is mixed with surface or
groundwater sources prior to drinking water treatment. Non-potable water reuse includes all
water reuse applications other than potable water reuse and it constitutes a large majority of
water recycling and reuse in the world.
266 T. Asano

4 THE ROLE OF WATER RECYCLING IN THE WATER CYCLE

The inclusion of planned water reclamation, recycling, and reuse in water resource systems
reflects the increasing scarcity of water sources to meet societal demands, technological
advancement, public acceptance, and improved understanding of public health risks. As the
link between wastewater, reclaimed water, and water recycling and reuse has become better
defined, increasingly smaller recycle loops are possible.
Traditionally, the hydrologic cycle has been used to represent the continuous transport of
water in the environment. The water cycle consists of fresh and salt water surface resources,
subsurface groundwater, water associated with various land use functions, and atmospheric
water vapor. Many sub-cycles to the hydrologic cycle exist including the engineered transport
of water such as canals and aqueducts. Water reclamation, recycling, and reuse, represent
significant components of the hydrologic cycle in urban, industrial, and agricultural areas.
A conceptual overview of the cycling of water from surface and groundwater resources to
water treatment facilities, irrigation, municipal, and industrial applications, and to wastewater
treatment and water reclamation/reuse facilities is shown in Figure 2. Water reuse may involve
a completely controlled pipe-to-pipe system with an intermittent storage step, or it may include
blending of reclaimed water with non-reclaimed water either directly in an engineered system
or indirectly through surface water supplies or groundwater recharge. The major pathways
of water reuse include groundwater recharge, irrigation, industrial use, and surface water
replenishment.
Surface water replenishment and groundwater recharge also occur through natural drainage
and through infiltration of irrigation and storm water runoff. The potential use of reclaimed
water for potable water treatment is also shown albeit rare and small quantity. The quantity
of water transferred via each pathway depends on the watershed characteristics, climatic and
geohydrological factors, the degree of water utilization for various purposes, and the degree
of direct or indirect water reuse.

Figure 2. The role of engineered wastewater treatment, reclamation, and reuse facilities in the cycling
of water as a sub-set of the hydrologic cycle (Adapted from Asano & Levine, 1998).
Water recycling – A relevant solution? 267

The water used or reused for agricultural and landscape irrigation includes agricultural,
residential, commercial, and municipal applications. Industrial reuse is a general category
encompassing water use for a diversity of industries that include power plants, pulp and paper
industries, and other industries with high rates of water utilization. In some cases closed-loop
recycle systems have been developed that treat water from a single process stream and recycle
the water back to the same process with some additional make-up water. In many industrialized
countries, zero liquid discharge regulations mandated water recycling and reuse within indus-
trial facilities. In other cases, reclaimed municipal wastewater is used for industrial purposes
such as in cooling towers. Closed-loop systems are also under evaluation for reclamation and
reuse of water during long-duration space missions by the National Aeronautics and Space
Administration (NASA).

4.1 The water quality changes


The water quality changes during municipal use of water in a time sequence is conceptually
shown in Figure 3.
Through the process of water treatment, a drinking water is produced which has an elevated
water quality meeting applicable standards for drinking water. The municipal and industrial
uses degrade water quality, and the quality changes necessary to upgrade the wastewater then
become a matter of concern of wastewater treatment. In the actual case, the treatment is
carried out to the point required by regulatory agencies for protection of other beneficial uses.
The dashed line in Figure 3 represents an increase in treated water quality as necessitated by
water reuse. Ultimately as the quality of treated water approaches that of unpolluted natural
water, the concept of water reclamation and reuse is generated. Further advanced wastewater
reclamation technologies, such as carbon adsorption, advanced oxidation, and reverse osmosis,
will generate much higher quality water than conventional drinking water, and it is termed
repurified water. Today, technically proven wastewater reclamation or purification processes
exist to provide water of almost any quality desired.

Figure 3. Water quality changes during municipal uses of water in a time sequence.
268 T. Asano

5 WATER RECLAMATION AND REUSE BENEFITS

Water reclamation and reuse make the best use of existing resources by: (1) conserving high-
quality water supplies by substituting reclaimed water for applications that do not require
that quality; (2) augmenting potable water sources and providing an alternative source of
supply to assist in meeting both present and future water needs; (3) protecting aquatic
ecosystems by decreasing the diversion of freshwater, reducing the quantity of nutrients
and other toxic contaminants entering waterways, and reducing the need for water control
structures; and (4) complying with environmental regulations and reuse liability by bet-
ter managing water consumption and wastewater discharges to meet regulatory limitations.
Benefits of water reclamation and reuse and factors driving its future are summarized in
Table 1.

Table 1. Benefits of water reclamation and reuse and factors driving its futurea .

Potential benefits of water reclamation and reuse


• Water recycling conserves water supplies:
Water recycling increases the total available water supply. High-quality water supplies can be
conserved by substituting reclaimed water where appropriate.
• Water recycling is environmentally responsible:
Water recycling can preserve the health of waterways, wetlands, flora and fauna. It can reduce
the level of nutrients and other pollutants entering water ways and sensitive marine environments
by reducing effluent and stormwater discharge.
• Water recycling makes economic sense:
Reclaimed water is at the door step of the urban development where water supply reliability is most
crucial and water is priced the highest.
• Water recycling can save resources:
Recycled water originating from treated effluent contains nutrients; if this water is used to irrigate
agricultural land, less fertilizer needs to be applied to the crops. By reducing pollution and nutrient
flows into waterways, tourism and fishing industries are also helped.

Factors driving its further use


• Increasing pressure on existing water resources due to population growth and increased agricultural
demand.
• Growing recognition among water and wastewater managers of the economic and environmental
benefits of using recycled water.
• Recognition that recycled water can be a reliable source of water supply even in drought
years.
• Increasing awareness of the environmental impacts associated with overuse of water supplies.
• Greater recognition of the environmental and economic costs of water storage facilities such as
dams and reservoirs.
• Preference to recycling over effluent disposal, coupled with tighter controls on the quality of any
effluent discharged to the environment.
• Community enthusiasm for the concept of water recycling.
• The growing numbers of successful water recycling projects in the world.
• The introduction of new water charging arrangements that better reflect the full cost of delivering
water to the consumers, and the widespread use of these charging arrangements.
• Increased costs associated with upgrading wastewater treatment facilities to meet higher quality water
quality standards.
a Compiled from various sources including Asano (1998), and Queensland Water Recycling Strategy (Queensland
Government, 2001).
Water recycling – A relevant solution? 269

6 WATER REUSE APPLICATIONS

To provide a framework for evaluating water reuse it is important to correlate major water
use patterns with potential reuse applications. On the basis of water quantity, agriculture
and landscape irrigation accounts for 54% of total freshwater withdrawals in the USA. The
second major user of water is industry, primarily for cooling and process needs. How-
ever, industrial uses vary greatly and additional wastewater treatment beyond secondary
treatment is usually required. Thus, the effective integration of water reuse into water
resource management is based on the quantity of water required for a specific applica-
tion and the associated water quality requirements. The major factors that influence water
reuse implementation are: (1) water quality requirements for intended water reuse applica-
tions; (2) existing or proposed wastewater treatment facilities; (3) requirements for degree of
treatment process reliability; (4) potential health risks mitigation; and (5) public perception
and acceptance.
Significant regional and seasonal variations in water use patterns exist. For example, in
urban areas, industrial, commercial and non-potable urban water requirements account for
the major water demand. In agricultural and arid zones, irrigation is dominant. Irrigation
requirements tend to vary seasonally whereas industrial water needs are more consistent. The
feasibility of water reuse for a given watershed depends on the practical extent that reclaimed
water could augment existing water supplies through substitution of water in commercial,
industrial, and agricultural applications.
The applications for municipal water reuse parallel major water use applications. An
overview of the seven major categories of water reuse is given in Table 2. These categories
are arranged in descending order of current and projected volume of reclaimed water with
applicable water quality requirements.
Much of the attention in recent years on the reclaimed water applications, however, has
been for its use in the urban environment, such as irrigation of parks and playgrounds and its
potential for groundwater recharge. It is interesting to note that Japan’s water recycling has
been decisively directed toward non-potable urban applications, such as toilet flushing, urban
environmental water, and industrial reuse. Comparison of water reuse in California and Japan
is shown in Figure 4. Only 16% of total water reuse in Japan (versus 68% in California) is for
agricultural and landscape irrigation.

Table 2. Categories of water reuse from treated municipal wastewater.

Category of
wastewater reuse Treatment goals Example applications

Urban use
unrestricted Secondary, filtration, disinfection Landscape irrigation: parks,
BOD5 ≤ 10 mg/L playgrounds, school yards,
Fecal coliform: ND/100 mL fire protection, construction,
Turbidity ≤ 2 NTU ornamental fountains, aesthetic
Cl2 residual: 1 mg/L impoundments.
pH: 6 to 9 In-building uses: toilet flushing,
air conditioning.
(Continued)
270 T. Asano

Table 2. (Continued)

Category of
wastewater reuse Treatment goals Example applications

restricted access Secondary and disinfection Irrigation of areas where public


irrigation BOD5 ≤ 30 mg/L access is infrequent and
TSS ≤ 30 mg/L controlled: freeway medians,
Fecal coliform ≤ 200/100 mL golf courses, cemeteries,
Cl2 residual: 1 mg/L residential, greenbelts.
pH: 6 to 9
Agricultural irrigation
food crops Secondary, filtration, disinfection Crops grown for human
BOD5 ≤ 10 mg/L consumption and consumed
Turbidity ≤ 2 NTU uncooked.
Fecal coliform: ND/100 mL
Cl2 residual: 1 mg/L
pH: 6 to 9
non-food crops Secondary, disinfection Fodder, fiber, seed crops,
and food crops BOD5 ≤ 30 mg/L pastures, commercial nurseries,
consumed after TSS ≤ 30 mg/L sod farms, commercial
processing Fecal coliform ≤ 200/100 mL aquaculture.
Cl2 residual: 1 mg/L
pH: 6 to 9
Recreational use
unrestricted Secondary, filtration, disinfection No limitations on body-contact:
BOD5 ≤ 10 mg/L lakes and ponds used for
Turbidity ≤ 2 NTU swimming, snowmaking.
Fecal coliform: ND/100 mL
Cl2 residual: 1 mg/L
pH: 6 to 9
restricted Secondary, disinfection Fishing, boating, and other
BOD5 ≤ 30 mg/L non-contact recreational
TSS ≤ 30 mg/L activities.
Fecal coliform ≤ 200/100 mL
Cl2 residual: 1 mg/L
pH: 6 to 9
Environmental reuse Site specific treatment levels Use of reclaimed wastewater to
pH create artificial wetlands,
Dissolved oxygen enhance natural wetlands and
Coliform, nutrients sustain stream flows.
Groundwater recharge Site specific Groundwater replenishment.
Salt water intrusion control.
Subsidence control.
Industrial reuse Secondary and disinfection Cooling-system make-up water,
BOD5 ≤ 30 mg/L process waters, boiler feed
TSS ≤ 30 mg/L water, construction activities
Fecal coliform ≤ 200/100 mL and washdown waters.
Potable reuse Safe drinking water requirements Blending in municipal water
supply reservoir.
Pipe to pipe supply.
Water recycling – A relevant solution? 271

Toilet
Environment
Environment
Agriculture
Agriculture
Landscape
Snow melt
Industrial
Industrial
Groundwater Cleansing

California: 434 Mm3 Japan: 206 Mm3

Figure 4. Comparison of water reuse in different categories in California and Japan.

7 OVERVIEW OF WASTEWATER TREATMENT TECHNOLOGY

The effective treatment of wastewater to meet water quality objectives for water reuse appli-
cations and to protect public health is a critical element of water reuse systems. Conventional
municipal wastewater treatment consists of a combination of physical, chemical, and biological
processes and operations to remove solids, organic matter, pathogens, and sometimes nutrients
from wastewater. General terms used to describe different degrees of treatment, in order of
increasing treatment level, are preliminary, primary, secondary, and advanced treatment. A
disinfection step for control of pathogenic organisms is often the final treatment step prior to
distribution or storage. Water reclamation and reuse treatment systems have developed from
technologies used for conventional wastewater treatment and drinking water treatment.
The goal in designing a water reclamation and reuse system is to develop an integrated
cost-effective treatment scheme that is capable of meeting water quality objectives reliably.

7.1 Water quality objectives


The contaminants in reclaimed water that are of public health significance include biological
and chemical constituents. Water quality control measures reference parameters that can be
classified as organic matter, pathogenic organisms, nutrients, toxic contaminants, or dissolved
minerals. Conventional wastewater treatment systems are typically designed to meet water
quality objectives based on biochemical oxygen demand (BOD5 ), total suspended solids (TSS),
total or fecal coliform, and turbidity.
Where reclaimed water is used for applications that have potential human exposure routes,
the major acute health risks are associated with exposure to biological pathogens including
bacterial pathogens, helminths, protozoa, and enteric viruses. From a public health and process
control perspective in developed countries, the most critical group of pathogenic organisms
is enteric viruses due to the possibility of infection from exposure to low doses and the lack
of routine, cost-effective methods for detection and quantification of viruses. In addition,
treatment systems that can remove enteric viruses effectively will most likely be effective for
control of other pathogenic organisms. Thus, it is essential to produce virtually virus-free
effluent for water reuse applications that have the potential for significant human exposure
or contact; e.g. spray irrigation of food crops eaten uncooked, parks and playgrounds, and
unrestricted recreational impoundments where swimming may take place.
272 T. Asano

The degree of treatment required in individual water treatment and wastewater reclam-
ation facilities varies according to the specific reuse application and associated water quality
requirements. The simplest treatment systems involve solid/liquid separation processes and
disinfection whereas more complex treatment systems involve combinations of physical,
chemical, and biological processes employing multiple barrier treatment approaches for con-
taminant removal. An overview of the major technologies that are appropriate for water
reclamation and reuse systems is shown in Figure 5. A summary of the major unit operations
and processes used in water reclamation is given in Table 3.

8 HEALTH AND REGULATORY REQUIREMENTS

In every water reclamation, recycling and reuse operation, there is some risk of human expos-
ure to pathogenic organisms and trace amount of harmful chemicals. Potential health impacts
are, however, related to the degree of human exposure to the reclaimed water and the adequacy,
effectiveness, and reliability of the treatment system. To protect public health without unneces-
sarily discouraging water reclamation and reuse, regulatory approaches stipulate water quality
standards in conjunction with requirements for treatment, sampling, and monitoring. To min-
imize health risks and esthetic problems, tight controls are imposed on the delivery and use
of reclaimed water after it leaves the treatment facility. Since major issues surrounding water
reclamation and reuse are often related to health, considerable amounts of research have been
conducted in the general area of public health. The major microbiological health hazards
associated with water consumption and contact with reclaimed water originate from fecal
contamination.
In the USA, comprehensive Federal standards for water reclamation and reuse do not exist
to date, and water reclamation criteria are developed by individual States, often in conjunction
with regulations on land treatment and disposal of wastewater. Some of the major differences
among the approaches taken by individual States are associated with the degree of specificity
provided in the rules. Also, discrepancies exist from place to place in terms of monitoring and
treatment requirements (Crook & Okun, 1987).
Reclaimed water quality criteria for protecting health in developing countries must be estab-
lished in relation to the limited resources available for public works and other health delivery
systems that may yield greater health benefits for the funds spent. Confined sewage collec-
tion systems and wastewater treatment are often non-existent in these countries and reclaimed
water often provides an essential water resource and fertilizer source. Thus, for most develop-
ing countries, the greatest concern for the use of wastewater for irrigation are caused by the
enteric helminths such as hookworm, Ascaris, Trichuris, and under certain circumstances, the
beef tapeworm. These pathogens can damage the health of both the general public consuming
the crops irrigated with untreated wastewater and sewage farm workers and their families.
The degree of treatment required and the extent of monitoring necessary depend on the
specific application. In general, irrigation systems are categorized according to the potential
degree of human exposure. A higher degree of treatment is required for irrigation of crops that
are consumed uncooked, or use of reclaimed water for irrigation of locations that are likely to
have frequent human contact.
To illustrate alternative regulatory practices governing the use of reclaimed wastewater
for irrigation, the major microbiological quality guidelines by the World Health Organiza-
tion (WHO, 1989) and the State of California’s (2001) current Water Recycling Criteria are
Low-Cost Wastewater Treatment
Irrigation/Reuse

Anaerobic Ponds Wastewater Storage and


Aerated Lagoons Maturation Ponds Treatment Reservoirs
Facultative Ponds Irrigation/Reuse
Upward-flow Anaerobic Sludge Bed (in series) for polishing and
pathogens removal

Conventional Wastewater Treatment


Preliminary Primary Secondary Tertiary/Advanced
Irrigation/Reuse

Irrigation/Reuse Irrigation/Reuse
Disinfection
Low-Rate Processes
Stabilization Ponds Disinfection Suspended Solids Removal
Disinfection Aerated Lagoons Chemical Coagulation
Filtration

Screening Nitrogen Removal


Communition High-Rate Processes
Sedimentation Nitrification – Denitrification
Grit Removal Activated Sludge Selective Ion Exchange
Trickling Filters Breakpoint Chlorination
Rotating Biological Contactors Gas Stripping
Overland flow
Secondary
Sedimentation Phosphorus Removal
Chemical Precipitation
Biological

Organics and Metals Removal


Sludge Processing Carbon Adsorption
Chemical Precipitation

Disposal/Reuse Dissolved Solids Removal


Reverse Osmosis
Electrodialysis
Distillation
Ion Exchange

Figure 5. Generalized wastewater treatment operations and processes, and water reclamation and reuse schemes (Pettygrove & Asano, 1985; Asano,
1998).
Water recycling – A relevant solution?
273
274 T. Asano

Table 3. Overview of unit operations and processes used in water reclamation.

Process Description Application

Solid/liquid separation
Coagulation Addition of chemicals to destabilize Promote particle destabilization to
suspended matter. improve flocculation and solids removal.
Flocculation Particle aggregation. Particle agglomeration upstream
of liquid/solid separation processes.
Filtration Particle removal by porous medium. Removal of particles larger than about
3 µM.
Sedimentation Gravity sedimentation of particulate Solids removal.
matter, chemical floc, and precipitates
from suspension by gravity settling.

Biological treatment
Aerobic biological Biological metabolism of waste solids Removal of organic matter from solution
treatment by bacteria in an aeration basin. by synthesis into microbial cells.
Oxidation pond Ponds with 2 to 3 feet of water depth Reduction of suspended solids,
for mixing and sunlight penetration. BOD, fecal bacteria and ammonia.
Disinfection The inactivation of pathogenic Protection of public health.
organisms using oxidizing chemicals,
ultraviolet light, caustic chemicals,
heat, or physical separation processes.

Advanced treatment
Activated carbon Process by which contaminants are Removal of hydrophobic organic
physically adsorbed onto the carbon compounds.
surface.
Air stripping Wastewater is distributed over a Used to remove ammonia nitrogen
packing through which forced air is and some volatile organics.
drawn to extract ammonia from the
water droplets.
Ion exchange Exchange of ions between an Softening and removal of selected ionic
exchange resin and water using contaminants; effective for removal of
a flow through reactor. cations such as calcium, magnesium,
iron and anions such as nitrate.
Lime treatment The use of lime to precipitate Used to stabilize lime-treated water,
cations and metals from solution. to reduce its scale forming potential.
Reverse osmosis Pressure membrane to separate ions Removal of dissolved salts from
from solution based on reversing solution.
osmotic pressure differentials.

compared in Table 4. The WHO guidelines emphasize that a series of stabilization ponds is
necessary to meet microbial water quality requirements. In contrast, the California criteria
stipulate conventional biological wastewater treatment followed by tertiary treatment including
filtration and chlorine disinfection to produce effluent that is virtually pathogen-free. Micro-
biological monitoring requirements also vary: the WHO guidelines also require monitoring
of intestinal nematodes, whereas the California Water Recycling Criteria rely on treatment
systems and monitoring of the total coliform density for assessment of microbiological quality.
Water recycling – A relevant solution? 275

Table 4. Comparison of microbiological quality guidelines and criteria for irrigation by the World
Health Organization (WHO, 1989) and the State of California’s (2001) current Water Recycling Criteria.

Intestinal Fecal or totalb Wastewater


Category Reuse conditions nematodesa coliforms treatment requirements

WHO Irrigation of crops <1/L <1000/100 mL A series of stabilization


likely to be eaten ponds or equivalent
uncooked, sports treatment.
fields, public parks.
WHO Landscape irrigation <1/L <200/100 mL Secondary treatment
where there is public followed by disinfection.
access, such as hotels.
California Spray and surface No standard <2.2/100 mLb Secondary treatment
irrigation of food recommended followed by filtration and
crops, high exposure disinfection.
landscape irrigation
such as parks.
WHO Irrigation of cereal <1/L No standard Stabilization ponds with
crops, industrial recommended 8–10 days retention or
crops, fodder crops, equivalent removal.
pasture, and trees.
California Irrigation of pasture No standard <23/100 mLb Secondary treatment
for milking animals, recommended followed by disinfection.
landscape impoundment.
a Intestinal nematodes (Ascaris and Trichuris species and hookworms) are expressed as the arithmetic mean number
of eggs per liter during the irrigation period.
b California Water Recycling Criteria is expressed as the median number of total coliforms per 100 mL, as determined
from the bacteriological results of the last 7 days for which analyses have been completed.

9 WATER REUSE PLANNING

Over the past decade, the impetus for water reuse has, in general, resulted from four motivating
factors: (1) availability of high quality effluents; (2) increasing cost of freshwater development;
(3) desirability of establishing comprehensive water resources planning, including water con-
servation and water reuse; and (4) avoidance of more stringent and expensive water pollution
control requirements such as needs for advanced wastewater treatment facilities (Asano &
Mills, 1990; Asano, 1991).
A common misconception in planning for water reuse is that reclaimed water repre-
sents a low-cost new water supply. This assumption is only true when water reclamation
facilities are conveniently located near large industrial or agricultural users, and when addi-
tional wastewater treatment is not required. The conveyance and distribution systems for
reclaimed water represent the principal cost of most water reuse projects. Recent experience
in California indicates that approximately US$ 2.50 in average capital costs are required for
making each 1 m3 /yr of reclaimed municipal wastewater available for water reuse. Assum-
ing a facilities life of 20 years and a 9% interest rate, the amortized cost of this reclaimed
water is in the neighborhood of US$ 0.24/m3 , excluding operation and maintenance (O & M)
costs.
276 T. Asano

The optimum water reuse project is best achieved by integrating both wastewater treatment
and water supply needs into planning. Thus, the facilities planning for water reuse should
consist of: (1) wastewater treatment and disposal needs assessment; (2) water supply and
demand assessment; (3) detailed reclaimed water market analysis; (4) engineering and eco-
nomic analyses of alternatives; and (5) implementation plan with financial analysis. These
planning steps are described in more detail in this section and important factors to consider in
planning and implementation are highlighted.

9.1 Planning basis


The typical framework for analysis is first to establish clearly defined objectives and identify
whether a project is intended as primarily single purpose or as multiple purpose, that is, to serve
two or more basic functions. Generally water reclamation projects serve the functions of either
water pollution control or water supply. Because most public works agencies, or subdivisions
of agencies, are established as single purpose entities, planning for water reclamation projects
is usually initiated with a single purpose in mind. For example, a city wastewater department
is confronted with the need to meet more stringent effluent discharge requirements and will
investigate water reclamation and reuse as one of the pollution control options. On the other
hand, a water department may be faced with a falling groundwater table and look upon water
reuse as a means of satisfying some of water demands by supplementing existing water supply.
In recent years, however, there are at least two simultaneous trends in the USA and other
countries which should be forcing us to view water reuse as fundamentally serving multiple
purposes: (1) standards for the discharge of wastewater are becoming increasingly more strin-
gent; and (2) freshwater resources are becoming increasingly stressed to meet growing and
competing water demands.
Many projects intended originally as single purpose inevitably have spillover benefits. If
these would be recognized at the outset of project planning, the scope could be expanded. By
recognizing their multiple benefits and beneficiaries, there are additional options available
in terms of sharing project responsibility and costs and achieving the optimum balance of
benefits (i.e. realizing maximum net benefits). The point of emphasizing the multiple purpose
concepts is that the traditional perspective of a single purpose agency and funding program is
often becoming outmoded and a disservice to meeting increasingly complex needs of society
with an environmentally conscious public.
The project study area is another critical planning issue. The typical approach is to equate
the study area with the project sponsor’s jurisdictional boundaries. However, this approach can
have serious pitfalls. The project study area should include all of the area that can potentially
benefit from reuse of effluent from a particular wastewater treatment plant. Because water
supply is typically dependent on water resources outside of the project study area, it is essential
to look beyond the local area to obtain an understanding of the water resources situation.
For example, over-drafted groundwater basins may be having their most serious impacts on
communities great distances beyond the local area. Thus, implementing water reuse in the
project area could result in a water supply savings in another area.
Planning for water reuse typically evolves through three stages: (1) conceptual level plan-
ning; (2) preliminary feasibility investigation; and (3) facilities planning. At the conceptual
level planning, a potential project is sketched out, rough costs estimated, and a potential
reclaimed water market identified. Based on the preliminary feasibility investigation, if water
Water recycling – A relevant solution? 277

reuse appears to be viable and desirable, then detailed planning can be pursued, refined
facilities alternatives developed, and a final facilities plan proposed.

9.2 Reclaimed water market assessment


A key task in planning a water reclamation project is to find potential customers that are
capable and willing to use reclaimed water. The approach to take in marketing the reclaimed
water depends on two factors: (1) project purpose – is the intent solely to treat and dispose of
the wastewater or also to obtain optimum water supply benefit? And (2) user option – will the
use of reclaimed water be voluntary or mandatory?
If the primary purpose is to treat and dispose of wastewater on land, then planners usually
seek land application sites where water can be applied at high rates, usually in excess of
optimum crop uptake rates, at least cost. Unless the system is designed with backup wastewater
disposal methods, the users will have to make a long-term commitment to accept the treated
effluent and may not have full control over the quantities of water delivered. If users cannot be
found to accept treated effluent on a voluntary contractual basis, the wastewater agency will
itself have to purchase wastewater application sites and apply the reclaimed water or lease the
land to a private farmer.
Projects designed with the primary purpose of water supply can usually be operated more
flexibly if alternate disposal, such as stream discharge, is available to dispose of effluent that
cannot be reused. The reclaimed water can be marketed on a voluntary basis. However, if
water supply is critical, the managing agency may elect to impose the use of reclaimed water
in place of freshwater where it is environmentally and humanly safe to do so.
Whether a user is capable of using reclaimed water depends on the quality of effluent
available and its suitability for the type of use involved. Willingness to use reclaimed water
depends on whether its use is voluntary and, if it is, on how well reclaimed water competes
with freshwater with respect to cost, quality, and convenience. It is essential to have a thorough
knowledge of the water supply situation, especially if reclaimed water is to be marketed on a
voluntary basis.

9.2.1 Market assessment


The water reuse market assessment consists of two parts: determination of background infor-
mation, and a survey of potential reclaimed water users and their needs. Important water
supply information includes a complete background on all of the wholesale and retail water
agencies in the planning area, their boundaries, quality of water served, prices charged, and
willingness to allow reclaimed water use in their jurisdictional areas. Because the introduction
of reclaimed water could reduce freshwater revenues, at least in the short run, there might be
resistance to implement wastewater reuse by some agencies. There should be the willingness
to consider the freshwater revenue impacts in the analysis and appropriate revenue and cost
sharing to obtain the full cooperation of all affected agencies.
Without much investigation it is possible to list most of the potential reclaimed water use
categories in the study area, e.g. landscape irrigation, industrial cooling, and irrigation of
food crops. On the basis of the use categories, health and water pollution control regulatory
authorities should be consulted to obtain their respective requirements. These would include
wastewater treatment requirements, on-site facilities modifications (e.g. backflow prevention
devices), and use area controls (e.g. no irrigation in areas of direct human contact). Technical
278 T. Asano

experts, such as farm advisors, can be consulted to determine acceptable water quality for
various use categories.

9.2.2 Identification of users


It is then possible to begin identifying and contacting individual potential users of reclaimed
water. Access to records of water retailers can be especially helpful. Several years of actual
water use records are helpful to ensure that planning is not misled by data from unusually wet
or dry precipitation years. It is important to obtain actual prices paid for water or, if a user has
its own supply, its fixed and variable costs. Potential users should be contacted and the reuse
sites visited to determine potential site problems or on-site water system modifications needed
to accommodate use of reclaimed water. These factors have cost implications which must be
assessed in the planning stage. The concern, needs, and financial expectations of users must be
identified. Group presentations with potential users may be useful to disseminate information
and make technical experts accessible to respond to questions.

9.3 Monetary analyses


While technical, environmental, and social factors are considered in project planning, mon-
etary factors tend to be overriding in the key decisions of whether and how to implement a
wastewater reuse project. Monetary analyses fall into two categories: economic analysis and
financial analysis, the distinction between which is critical. The economic analysis focuses
on the value of the resources invested in a project to construct and operate it, measured in
monetary terms and computed in the present value. On the other hand, the financial analysis
is based on the market value of goods and services at the time of sale, incorporating any par-
ticular subsidies or monetary transfers which may exist. Whereas economic analysis evaluates
wastewater reuse projects in the context of impacts on society, financial analysis focuses,
instead, on the local ability to raise money from project revenues, government grants, loans,
and bonds to pay for the project.
The basic result of the economic analysis is to answer the question, should a reuse project be
constructed? Equally important, however, is the question, can a reuse project be constructed?
Both orientations, therefore, are necessary. However, only water reuse projects which are
viable in the economic context should be given further consideration for a financial analysis
(Mills & Asano, 1998).

9.3.1 Economic analysis


The role of an economic analysis is to provide a basis for justifying a wastewater reuse project
in monetary terms. A project is considered justified if its total benefits exceed its total costs. If
several alternatives can meet the same objective, then the alternative providing the maximum
net benefit is the economically justifiable project. While the benefit-to-cost ratio is a com-
mon measure of economic justification, it is not the best measure to determine the optimum
project size.
An important aspect of the economic analysis is that it takes into consideration all costs
and benefits associated with the alternatives under consideration, placing all alternatives on
equal footing for comparison. Also, this analysis is completely independent of financing
considerations. To identify all costs and benefits it is essential to look beyond the boundaries
of the agency doing the planning. For example, an agency may be seeking a new source of
Water recycling – A relevant solution? 279

water supply from outside of its boundaries. To perform an economic comparison it would be
necessary to identify the construction, operation, and maintenance costs of this supply.

9.3.2 Cost and price of water


Another important aspect of an economic analysis is that it considers only the future flow of
resources invested in or derived from a project. Past resources investments are considered sunk
costs that are irrelevant to future investment decisions. Thus, debt service on past investments
is not included in an economic analysis. A common error in this respect is to confuse water
price with water cost. Water price is the purchase price paid to a water wholesaler or retailer to
purchase water and usually reflects a melding of current and past expenditures for a combin-
ation of projects, as well as water system administration costs, which are generally fixed costs.
The costs of relevance to an economic analysis would be only costs for future construction,
operations, and maintenance. If a water reuse project were to be compared to a particular
new water supply development, the relevant costs would be the future stream of costs: (1) to
construct new freshwater facilities; and (2) to operate and maintain all of the facilities needed
to treat and deliver the new increment of water supply developed. This stream of costs may
bear no resemblance to the present and future price, at the wholesale or retail level, charged
for water.
In contrast, water prices embody debt service on existing facilities, and future projections
are an average price to recover costs for both existing facilities and future additions. Typically,
water price will be much lower than the marginal cost of developing a new water supply,
because the cost of each new source of supply is increasingly expensive due to inflation and
the greater difficulty in developing new supplies.

9.3.3 Financial analysis


The financial analysis addresses the question whether a wastewater reuse project is financially
feasible. The project sponsor will need a source of capital and sources of revenue to pay
for debt service and operational costs for both the proposed reuse project and any existing
facilities. Fixed costs for existing facilities, while irrelevant in the economic analysis, must be
considered in a financial analysis if they are a continuing financial obligation.
The water reuse project sponsor is not the only important party to consider in a financial
analysis. Of particular importance is the participation of the user of the reclaimed water.
The user will be expecting a net cost of reclaimed water that is no more than it would have
paid for freshwater. For example, a reclaimed water customer may have to invest in piping
modifications or a dual water supply system to accommodate the reclaimed water. On the
other hand, a farmer may be able to save on fertilizer costs by taking advantage of nitrogen
and phosphorus contained in reclaimed water. A prospective user will expect the difference in
price between freshwater and reclaimed water to reflect any added costs or savings.
Because the sale of reclaimed water may reduce revenue from freshwater sale, there may
be a need to evaluate the effect on the freshwater retailer and freshwater prices. It may be
necessary to allocate some of the reclaimed water revenue to compensate for the freshwater
revenue loss. On the other hand, if reclaimed water offsets the purchase or development of
more expensive freshwater, then it may be appropriate for freshwater revenue to be used to
subsidize the wastewater reuse project.
It is not uncommon that potential users may have different sources of water or have different
rate schedules. It is important, therefore, not to assume that there is an average price that all
280 T. Asano

users are paying for freshwater. Failure to take into account the financial situation of each
user could result in the loss of key reclaimed water customers. The initial market assessment
should have included this financial data. In conclusion, there should be flexibility in tailoring
a financial scheme to fit each situation best.

9.4 Other planning factors


A number of factors besides the monetary aspects have to be evaluated during the planning for
a water reuse project, such as environmental impacts. However, factors of particular signifi-
cance in project development are related to engineering and public health. Engineering involves
more than water distribution system design. A water reuse project is a relatively small-scale
water supply project with considerations of matching supply and demand, appropriate levels
of wastewater treatment, reclaimed water storage, and supplemental or backup freshwater
supply.

9.4.1 Water demand characteristics


In freshwater systems, water demands are first projected and then water supplies are developed
to meet the demand. The reverse procedure is often applied in water reuse system planning.
The wastewater supply rate is accepted as a given and the reclaimed water demand is added
to the system until the economic optimum is met. For example, landscape irrigation demand
in California is seasonal. However, wastewater production is nearly constant year-round.
Reclaimed water supply may be sufficient to meet annual demands, but only if seasonal storage
is available. Seasonal storage, however, is costly and, in urban settings, even impossible to site.
Another option is to include fewer users in the system such that the peak demands can be
met entirely by the reclaimed water supply without seasonal storage. This could, however,
result in the waste of as much as 40% of the available reclaimed water. What will probably
be the optimum situation is to add users in excess of supply and meet the peak demands with
supplemental freshwater. There may still be some supply that cannot be used or economically
stored during low water demand periods, but this lost supply can be reduced substantially
because of the availability of a supplemental supply in the peak season. Some projects have
incorporated an added benefit by utilizing a poor quality water supply unsuitable for potable
use, such as an abandoned groundwater basin, to supplement reclaimed water.

9.4.2 Supplemental water supply


Supplemental freshwater can be blended with reclaimed water in the distribution system or on
the user’s site. Because of public health concerns about potential cross connection or backflow
of reclaimed water into potable water supply systems, it may be necessary to provide an air
gap between the supplemental supply and the reclaimed system.
Even if supplemental water is not needed to meet demands, there may still be a need to pro-
vide an emergency backup water supply during periods of treatment plant upset or equipment
failure. Because a backup water supply would be utilized in place of, rather than simultan-
eously with, reclaimed water, there are more options for introducing it into the system. With
appropriate backflow prevention the reclaimed water distribution system can be connected to
the potable system during the emergency period. It should be noted that with the availability
of a backup water supply, there is less need for equipment redundancy in the reclaimed water
system to ensure 100% reliability.
Water recycling – A relevant solution? 281

9.4.3 Water quality


If a significant market could be added by upgrading reclaimed water quality, project alterna-
tives should be developed for various treatment levels. The levels of wastewater treatment and
water quality for landscape and agricultural irrigation uses are normally governed by health-
related regulations, though crop sensitivity to effluent constituents such as salts and boron
should be investigated (Ayers & Westcot, 1985; Pettygrove & Asano, 1985).
Industrial users will have more stringent physical and chemical water quality requirements
that will affect levels for wastewater treatment. With some exceptions, it is impracticable to
serve more than one quality of water. Thus, the level of wastewater treatment provided may
be higher than many of the users actually require. If there is a reclaimed water market for
two levels of water quality, it should be considered whether the distribution system can be
separated so that the higher and more expensive treatment can be sized to serve only those
users needing such higher water quality.
While the emphasis on the wastewater disposal or water supply aspects will vary depending
on whether a project is single or multiple purposes, the nature of water reuse is such that both
aspects must at least be considered. Even if it is determined that a water reuse project is not
feasible at the conclusion of the study, it is still advisable to publish the information and data
collected and the analysis performed to arrive at this conclusion. Water reuse is good public
policies in appropriate situations and the public interest in them will continue to recur as long
as water supply needs are perceived to be critical, such as in drought years. Documentation of
even unsuccessful reuse planning is helpful in responding to public inquiry and in orienting
future planning efforts.

10 WATER REUSE: A RELEVANT SOLUTION

Inadequate water supply and water quality deterioration represent serious contemporary con-
cerns for many municipalities, industries, agriculture, and the environment in various parts of
the world. Several factors have contributed to these problems such as continued population
growth in urban areas, contamination of surface water and groundwater, uneven distribution of
water resources, and frequent droughts caused by the extreme global weather patterns. For more
than a quarter century, a recurring thesis in environmental and water resources engineering
has been that improved wastewater treatment provides a treated effluent of such quality that it
should not be wasted but put to beneficial use. This conviction in responsible engineering, cou-
pled with the vexing problem of increasing water shortage and environmental pollution, pro-
vides a realistic framework for considering reclaimed water as a water resource in many parts
of the world. Water reuse has been dubbed as the greatest challenge of the 21st century as water
supplies remain practically the same and water demands increase because of increasing popula-
tion and per capita consumption. Water reuse accomplishes two fundamental functions: (1) the
treated effluent is used as a water resource for beneficial purposes; and (2) the effluent is kept
out of streams, lakes, and beaches; thus, reducing pollution of surface water and groundwater.

10.1 The main concerns in water reuse


The main concerns in water reuse are: (1) reliable treatment of wastewater to meet strict water
quality requirements for the intended reuse; (2) protection of public health; and (3) gaining
282 T. Asano

public acceptance. Water reuse also requires close examinations of infrastructure and facilities
planning, and water utility management involving effective integration of water and reclaimed
water supply functions. Whether water reclamation and reuse will be appropriate in a specific
locale depends upon careful economic considerations, potential uses for the reclaimed water,
and stringency of waste discharge requirements. Public policy wherein the desire to conserve
and reuse rather than develop additional water resources with considerable environmental
expenditures may be an important consideration. Through integrated water resources planning,
the use of reclaimed water may provide sufficient flexibility to allow a water agency to respond
to short-term needs as well as increase long-term water supply reliability.

10.2 Groundwater recharge and direct potable water reuse


Groundwater recharge with reclaimed water and direct potable water reuse share many of
the public health concerns encountered in drinking water withdrawn from polluted rivers and
reservoirs. Three classes of constituents are of special concern where reclaimed water is used in
such applications: (1) enteric viruses and other emerging pathogens; (2) organic constituents
including industrial chemicals, home care products, and medicines; and (3) heavy metals.
The ramification of many of these constituents in trace quantity are not well understood with
respect to long-term health effects, and, as a result, regulatory agencies are proceeding with
extreme caution in permitting water reuse applications that affect potable water supplies. In
all the cases in the USA where potable water reuse has been contemplated, alternative sources
of water have been developed in the ensuing years and the need to adopt direct potable water
reuse has been avoided. As the proportional quantities of treated wastewater discharged into
the nation’s waters increase, much of the research which addresses groundwater recharge and
potable water reuse is becoming of equal relevance to unplanned indirect potable reuse such as
municipal water intakes located downstream from wastewater discharges or from increasingly
polluted rivers and reservoirs.

10.3 Locally controllable water resource


Reclaimed water is a locally controllable water resource that exists right at the doorstep of the
urban environment, where water is needed the most and priced the highest. Closing the water
cycle not only is technically feasible in industries and municipalities but also makes economic
sense. While potable reuse is still a distant possibility and may never be implemented in
the USA, groundwater recharge with advanced wastewater treatment technologies is a viable
option backed by the decades of experiences in Arizona, California, New York, and Texas, as
well as in Australia, Israel, Germany, the Netherlands, and the UK. It is important to recognize
that public acceptance of water reuse projects is vital to the future of water reclamation,
recycling, and reuse; the consequences of poor public perception could jeopardize future
projects involving the use of reclaimed water. Indeed, water recycling and reuse provide a
relevant solution to water resources development and management but with more challenges
in costs, health protection, and public acceptance.

REFERENCES

Angelakis, A.N. & Spyridakis, S.V. (1995). The Status of Water Resources in Minoan Times: A Prelim-
inary Study. In: A.N. Angelakis, A. Issar & O.K. Davis (eds.), Diachronic Climatic Impacts on Water
Resources in Mediterranean Region. Springer-Verlag, Heidelberg, Germany.
Water recycling – A relevant solution? 283

Asano, T. (1991). Planning and Implementation of Water Reuse Projects. Water Science and Technology,
24(9): 1–10.
Asano, T. (ed.) (1998). Wastewater Reclamation and Reuse. Water Quality Management Library. Vol.
10, CRC Press. Boca Raton, Florida, USA.
Asano, T. & Levine, A.D. (1996). Wastewater Reclamation, Recycling, and Reuse: Past, Present and
Future. Water Science and Technology, 33(10–11): 1–14.
Asano, T. & Levine, A.D. (1998). Wastewater Reclamation, Recycling, and Reuse: An Introduction, In
Asano, T. (ed.), Wastewater Reclamation and Reuse. Water Quality Management Library, CRC Press.
Boca Raton, Florida, USA.
Asano, T. & Mills, R.A. (1990). Planning and Analysis for Water Reuse Projects. Journal American
Water Works Association, January: 38–47.
Ayers, R.S. & Westcot, D.W. (1985). Water Quality for Agriculture. FAO Irrigation and Drainage Paper
29, Rev. 1. Food and Agriculture Organization of the United Nations, Rome, Italy.
Barty-King, H. (1992). Water the Book, an Illustrated History of Water Supply and Wastewater in the
United Kingdom. Quiller Press Limited, London, UK.
Crook, J. & Okun, D.A. (1987). The Place of Non-potable Reuse in Water Management. Jour. Water
Pollution Control Fed., 59(5): 236.
EEC (European Communities Commission) (1991). Council Directive regarding the treatment of urban
wastewater (91/271/EEC). Official Journal of the European Communities, No. L 135, 30 May: 40–50.
Fair, G.M. & Geyer, J.C. (1954). Water Supply and Wastewater Disposal. John Wiley and Sons,
New York, USA.
Levine, A.D. & Asano, T. (2004). Recovering Sustainable Water from Wastewater. Environmental
Science & Technology, 38(11): 201A–208A.
Mills, R.A. & Asano, T. (1998). Planning and Analysis of Water Reuse Projects. In Asano, T. (ed.),
Wastewater Reclamation and Reuse. Water Quality Management Library. Vol. 10, CRC Press, Boca
Raton, Florida, USA.
Pettygrove, G.S. & Asano, T. (eds.) (1985). Irrigation with Reclaimed MunicipalWastewater: A Guidance
Manual. Lewis Publishers, Inc., Chelsea, Michigan, USA.
Postel, S.L. (2000). Entering an era of water scarcity: the challenges ahead. Ecological Applications,
10(4): 941–948.
Queensland Government (2001). Queensland Water Recycling Strategy: An Initiative of the Queensland
Government. The State of Queensland, Environmental Protection Agency, Queensland, Australia.
State of California (2001). Water Recycling Criteria, an excerpt from the California Code of Regulations,
Title 22, Division 4. Environmental Health, Department of Health Services, Sacramento, California,
USA.
United Nations (2003). World Population Prospects: the 2002 Revision – Highlights. United Nations
Population Division, Department of Economic and Social Affairs, United Nations, New York, USA.
WHO (World Health Organization) (1989). Health Guidelines for the Use of Wastewater in Agriculture
and Aquaculture, Report of a WHO Scientific Group. Geneva, Switzerland.
Young, D.D. (1985). Reuse via rivers for water supply, In: Reuse of Sewage Effluent, Proceedings of
the International Symposium organized by the Institution of Civil Engineers. London, 30–31 October
1984. Thomas Telford, London, UK.
CHAPTER 17

Urban and industrial watersheds and ecological


sanitation: two sustainable strategies for on-site urban
water management

D. Del Porto
Ecological Engineering Group, Concord, Massachusetts, USA

ABSTRACT: Many of the water crises in urban areas are based not so much on the availability of
water but rather on our failure to manage how water is used. This is due to the failure to recognize
stormwater as a water source and using water to transport unwanted wastes. Wastewater recycling is
an important part of an integrated water strategy – but not the final answer. That integrated strategy
includes stormwater collection, source separation, wastewater and graywater recycling for potable and
non-potable use, and advanced conservation.

Keywords: Integrated water planning strategy, stormwater, ecological sanitation, wastewater,


watershed, recycling, reuse, conservation

1 INTRODUCTION

“Problems cannot be solved at the same level of awareness that created them”.
(Albert Einstein)

1.1 The urban dilemma


As the world’s population increases, migrates to coastal areas and builds new cities, urban plan-
ners are confronted with insufficient supplies of fresh water to meet the increasing demands.
As a consequence, supply augmentation, water demand management practices, and water
recycling are expected to supplement the gap between available supply and new demand. The
problem is not limited to arid zones. It is a function of population explosions and the inability
to manage demand. “Even in water-rich countries such as Canada, lack of quality control
for drinking water has led to death and illness in several communities, and some provinces
suffer chronic shortages of water for agriculture; and almost everywhere capital costs for
infrastructures to supply and remove water are growing” (Brooks, 2003).
Much of the water crises in urban areas are based not so much on the availability of water but
rather on our failure to manage how it is used. Both the failure of urban planners to recognize
stormwater as a resource and the continued use of water to transport unwanted wastes are at
the root of the problem.

1.2 Problems with the current approach


Addressing urban water supplies is quickly becoming a high-priority challenge. The United
Nations Population Fund estimates that by 2025, 61% of the world’s population of 8000 million
286 D. Del Porto

will live in cities (United Nations, 2004), creating population densities throughout the world
heretofore unseen.
The problem is exacerbated as urban development follows the western model of big pipes
bringing fresh water from long distances and then treating influent wastewater and discharg-
ing into receiving waters. Or, it flows to sophisticated expensive recycling treatment plants
and then distributed through more big pipes to where the water was wasted in the first
place.
Continuing this model, urban planners will have to seek:
– New supplies from other aquifers and surface waters (yet these resources are already
overtaxed in many areas and may rob water from outlying areas and agriculture).
– Desalinization (very energy intensive).
– Wastewater recycling (energy and infrastructure intensive).
It is clear that wastewater recycling will be a necessary strategy in the years ahead. However,
it is only one component of an ecological and integrated strategy that makes best use of local
sites as well as local sources of water and local effluents that can be reused.
Continuing to combine wastes, including excreta, toxics, and heavy metals, in wastewater
flows, then turning to advanced ultra-filtration at the end of the pipe will not be feasible for
many cities, especially those in developing countries.
To identify water recycling as the best solution to the world’s water supply challenges is to
use the same mentality that created today’s many water problems.
The practice of centrally collecting combined effluents from a wide variety of sources
containing many constituents, and then treating it with end-of-pipe solutions has evolved in
an effort to avoid the complexity of using many smaller, more local and effluent-specific
strategies. Yet nature’s model shows us that complexity is the best way to manage resources.
For this reason, urban wastewater recycling can only be viewed as an important part of
a much larger picture of integrated water planning. Such a strategy, which decentralizes
water collection, use and treatment, nevertheless still benefits from central management.
Also, in contrast to the big pipe approach, this strategy requires us to heed the tenet start
at the source. By first seeking more local opportunities – both on the site of the water use
and wastewater discharge and according to what exactly is discharged – we may find eco-
logical and low-entropy possibilities for reducing water demand and treating effluents in
more ecologically effective ways. We will call this integrated source- and site-based water
planning.

1.2.1 Untapped water sources


One aspect of integrated source – and site-based water planning is identifying all of the water
sources available to a particular site of water use.
Consider the following example: on average, New York City receives about 937 Mm3 /yr
(246,500 million gal/yr) of stormwater, or 2.57 Mm3 (675 million gal) per averaged day
(New York City has 836.1 km2 of land area and receives 122.6 mm/yr of total precipitation).
While some is lost through evaporation, most is piped away into the adjacent rivers and
out to sea.
New York City uses about 4.16 Mm3 /d (1100 million gal/d). According to the New York
City Water Department, the city saved 1.5 Mm3 (400 million gal) of water by replacing 1.3
million old water-wasting toilets, installing locks on fire hydrants in neighborhoods throughout
Urban and industrial watersheds and ecological sanitation 287

the five boroughs, and implementing a leak detection program to inspect underground water
mains for leaks.
Therefore, if the stormwater had been collected, it would have satisfied 61.4% of the
current demand. Using the current price: US$ 1.39 per m3 (3.94 per 100 ft3 ) (New York City
Water Department) for water service, the value of the stormwater is US$ 3.56 million per day
or US$ 1300 million per year.
Stormwater could satisfy 100% of the demand if: (1) demand is reduced by an additional
22% by continued conservation; and (2) recycled wastewater supplied 20% of the demand.
Yet the city draws its water from various watersheds in distant mountains, requiring miles
of pipes and straining the communities along the watershed. New York City might instead
look to capturing this stormwater, using some of the thousands of millions of dollars spent
for collection and distribution pipe to fund cisterns, rooftop collection, and simple rainwater
treatment to augment water supply.

1.3 Recycled wastewater


While recycling wastewater is an important step in the conservation of fresh water, it is
technologically complex, expensive, and unacceptable to many communities, such as San
Diego, California, according to the USA-based WateReuse Association (Web Page). In addi-
tion, there is a growing concern for the impacts of recycled water on valuable aquifers
due to contamination by constituents that are currently not monitored by groundwater dis-
charge permits. According to the California Department of Health Services (Web Page),
these aquifer anti-degradation concerns are reported as action levels. Action levels are health-
based advisory levels for chemicals in drinking water that lack maximum contaminant levels
(MCLs).

1.4 The QWERTY syndrome


We humans have a curious tendency to continue old practices into the future long after the
need for such practices has passed. Take the example of typing keyboards. The upper left
row of letters begins with the letters QWERTY. This arrangement was developed in 1872 to
slow down typists; otherwise the keys would jam on the early mechanical typewriters. We no
longer have mechanical typewriters, yet we continue to use the old keyboard configurations
to this day. It may be that this QWERTY syndrome is the root cause of some of today’s most
fundamental water management problems.
There are two QWERTY -related water issues that will be the focus of this chapter:

– Urban and industrial stormwater management attitudes: we must reverse the attitude that
stormwater in an urban and industrial setting is not a resource, but rather a problem due to
flooding and so we must quickly pipe it away.
– Historical linkage of water and sanitation: one of the main impediments to sustainable water
management is the historic linkage of water and sanitation. We must uncouple water from
sanitation as much as possible or forever be trapped in flushing away unwanted residuals,
valuable nutrients (found in urine), and pathogens with large volumes of water. An example
is New York City’s expensive flush. As of the 2000 census, the population of New York
City was 8.1 million. Assuming that the average person flushes a toilet 5.1 times per day
(Vickers, 2001) and that the average water use per flush is 11.36 L (3 gal), then New York
288 D. Del Porto

City flushes about 0.47 Mm3 (123.75 million gal) worth US$ 0.65 million of drinking water
into the sewers every day.

1.4.1 The fundamental problem


Essentially, we are using a valuable resource – drinking water treated and delivered at signifi-
cant expense – to dilute and dispose of another potentially valuable resource, human excreta.
To this we add industrial and household chemicals and stormwater drainage. Then we pay a
very high cost to transport this combined effluent to a facility which attempts to separate all
of those constituents, clean them to a degree mandated by national law, and discharge the
remaining water back into the environment – usually rivers, oceans, and the ground. In most
cases, the same nutrients and toxic chemicals that went into the wastewater mix are still present
in what leaves the treatment plant. Growing realization of the effects of this is prompting regu-
lators to mandate further treatment – and that is making this combine-dilute-treat-and-dispose
approach very expensive.
However, even with ever-tightening requirements for advanced wastewater treatment, many
of the world’s waterways do not meet the goals of swimmable and fishable quality set by the
USA 1972 Clean Water Act nor the United Nation’s Agenda 21 (Brooks, 2003). And, while
much progress has been made over the last 32 years, there are many who ask if there might
not be a better way.
Every year we are learning more about the longer-term effects of partial treatment and
disposal, our present approach to cleaning wastewater. In USA, the 1972 Clean Water Act
only required the reduction of suspended solids, biological oxygen demand and fecal coli-
form bacteria to protect lakes and rivers. But now, responsible regulators worldwide are
mandating the removal of nutrients, toxic chemicals, parasites, viruses, radioactive wastes
and other constituents. At the same time, planners are asking: in a world where drinking
water is increasingly expensive and scarce, can we use this valuable resource for flushing
toilets?
It is clear that better ways are needed; the good news is that many are here and more are
emerging. However, the answer is not merely a matter of more clean-up at the end of the
sewage pipe. A larger, more strategic solution is called for based on a broad approach:

– Advanced water conservation.


– Collect and store stormwater.
– On-site wastewater treatment.
– Recycling.

2 BETTER WAYS EMERGING: FROM DISPOSAL TO UTILIZATION

More regulations requiring better treatment, and increasing costs for water and wastewater
treatment are prompting a reframing of the wastewater issue. In the current system, we create
wastes that we want disposed of. A better strategy is to put these outputs to use, just as they
are in nature’s model. In balanced ecosystems there is no waste: the outputs of one organism
are the inputs of another.
The solution to both of these problems is the development of a new ecological paradigm
for integrated water management. To accomplish this integration we must investigate the new
Urban and industrial watersheds and ecological sanitation 289

initiatives that address these issues. Key initiatives can be summarized as urban and industrial
watersheds and ecological sanitation.

2.1 Urban and industrial watersheds


The ecological engineering approach for water management presented in this report will not
only demonstrate environmental sustainability for the 21st century, but will also be testaments
to these visions. The principles of industrial ecological and sustainable development do not
preclude economic development or manufacturing efficiency. Rather, recycling, reuse and con-
servation result in reduced dependence on scarce resources, higher efficiency in manufacturing
processes, and long-term cost savings.
Terrestrial life efficiently self-organizes around water and the mineral-rich landforms
through which water flows. The ecosystem that best describes life and all related activities
within the water and landform context is a watershed. A natural watershed synthesizes inputs
of rainwater, solar energy and minerals from within its physical, chemical, and biotic com-
munities to produce an array of nutrients, raw materials and products that sustains a certain
quality of life for all its inhabitants. Concerning society’s welfare, almost all essential land-
scape functions (i.e. fertile soil regeneration, climate stabilization, etc.) for are connected to
water ecology. In the same way, the industrial watershed is defined by its rainwater, wastewater
and process water flows and its interaction with the surrounding watershed.

2.1.1 Ford Motor Company Rouge Facility as an Industrial Watershed


(Del Porto & The Ecological Engineering Group, 2001)
According to Environment Canada (Web Page), “an automobile coming off the assembly
line, for example, will have used at least 80,000–120,000 L (21,000–32,000 gal) of water to
produce its ton of steel and 40,000 L (10,600 gal) more for the actual fabrication process.
Many thousands more liters of water are involved in the manufacture of its plastic, glass,
fabric components”.
The Ford Rouge Plant in Dearborn, Michigan – with its 243 ha (600 acres) of natural and
constructed landforms, inputs of water, energy, minerals, capital, information, and nutrients
from its 30,000 employees, and outputs of products that sustain the Dearborn community – is
in fact a constructed ecosystem that we will call an Industrial Watershed.

2.1.2 A water eco-nomic example


With increasing prices and competition, water may arise as an important cost factor within the
production process. Therefore, new water management strategies, not only in conservation
but also in terms of reuse, are needed. Following figures show what the gain of water from
the industrial watershed could mean to the production of automobiles.

2.1.3 Source: stormwater


About 2 Mm3 /yr (530 million gal/yr) of water falls on the 243 ha (600 acres) Rouge Plant
site and discharges to the Rouge River. Average precipitation is 825.5 mm/yr (32.5 inches/yr)
based on 30-year average (NRCS/US Department of Agriculture weather data). Increasingly
this water is coming under state and federal jurisdiction as stormwater discharge controls are
implemented to protect the Rouge River. The federal government, in order to limit pollutants in
stormwater discharges from industrial facilities, implemented the National Pollutant Discharge
290 D. Del Porto

Elimination System (NPDES), Phase I Storm Water Program, which includes an industrial
stormwater permitting component. The requirement for a NPDES industrial stormwater permit
has added an additional financial component to the Ford water management costs. While the
Michigan counterpart is voluntary, one can predict mandatory compliance will be coming in
the near future.

2.1.4 Application: process water


General Motors de Mexico (Ramos Arizpe Complex) recently received the Stockholm Industry
Water Award for demonstrating sustainable water and wastewater treatment and recycling tech-
niques. Using a variety of physical, chemical and biological wastewater treatment processes,
the facility was able to recover and reuse 70% of its industrial wastewater. The facility also
increased its production seven-fold, while reducing the average amount of well water needed
to produce a vehicle from 32 m3 (8452 gal) to 2.2 m3 (581 gal).
The success of General Motors may serve as an incentive and model for other industrial
operations that seek to minimize environmental impact while maximizing production and
quality of life.
If Ford, using General Motors water usage, would collect 90% of the stormwater that
would provide 1.8 Mm3 (477 million gal) which could be used to make 820,998 vehicles
(477 × 106 /581) without taking any from the Dearborn Water Company. Because the same
water can be treated (in respect to its constituents) and reused with about 90% recovery effi-
ciency, another 738,898 automobiles could be made (429.3 × 106 /581) after treatment. The
remaining treated water could be sold back to Ford Steel or the Dearborn Water Department
for a profit or used to grow some of the raw materials. Therefore, 1,559,896 vehicles per
year could be made with only one recycle pass and still have recycled water to use for other
profitable purposes!

2.1.5 Source: waste water


In 2001, the price of water service from the Dearborn Water Department via the Detroit Water
and Sewer Department was comprised of the cost to treat and deliver potable water from Lake
Huron and the Detroit River, as well as the cost of collecting and treating it as wastewater once
it has been used (see Detroit News Online, Web Page). According to Dearborn, Michigan’s
1999 annual report, the city spent US$ 25.3 million on the city’s water distribution and sewage
collection systems. Of the total bill, wastewater management is more than 66% of the total.
Upgrading the aged infrastructure will cost US$ 5300 million, which in the short-term will
increase the sewage portion of the total bill by 13.7%.
The 2001 unit cost for water of US$ 0.87 per m3 (US$ 3.30 per 1000 gal) does not reflect
the cost of infrastructure upgrade. Therefore, the unit cost does not reflect all the new marginal
costs and will certainly escalate dramatically in the future.

2.2 Water-wise design through ecological engineering


The management of the industrial watershed to sustain the highest quality of life for the low-
est cost of living involves adopting an engineering model that is founded on the ecological
paradigm. The transformation of an industrial site into a highly efficient and therefore prof-
itable industrial watershed will require the ecological engineering of all aspects of the facility.
Ecological engineering is predicated on the knowledge that the self-organizing order found
in stable ecosystems is so universal that it can be applied as an engineering discipline to
Urban and industrial watersheds and ecological sanitation 291

solve the pressing problems of global pollution, goods and services production, and efficient
resource-utilization, while providing a high quality of life.
When possible, we must recycle, reuse or utilize effluents. Using them strategically, such
as for landscape irrigation, flushing toilets, and evaporative cooling, helps save both water
supply and wastewater treatment costs – and prevents effluents from becoming pollution.
Just as reduce, reuse and recycle has become the credo of responsible solid waste man-
agement, this five-pronged strategy will become more obvious and important to the world’s
population than presently can be imagined.
Increasing costs of water and wastewater treatment infrastructure are driving the interest
for integrated water management, by viewing urban, domestic, commercial, industrial and
agricultural landscapes and hardscapes as watersheds that collect water, use water, and dispose
of stormwater and wastewater within larger regional watersheds.

2.3 Ecological sanitation


According to the Deutsche Gesellschaft für Technische Zusammenarbeit (GTZ, Web Page),
“Conventional forms of centralized sanitation are coming under increasing criticism. Espe-
cially because of the enormous investment involved, the huge operating and maintenance costs,
high water consumption and other drawbacks, they are not suitable as a blanket solution for
developing countries, particularly in arid climate zones. Even conventional individual disposal
systems, such as latrines and cesspits, make poor alternatives – especially in view of increas-
ing population densities and the substantial groundwater pollution they cause. Moreover, all
conventional types of wastewater and sewage disposal systems usually deprive agriculture,
and consequently food production, of the valuable nutrients contained in human excrement.
A more holistic approach towards sustainable sanitation is offered by the concepts referred
to as ecological sanitation. The key objective of this approach is not to promote a certain
technology, but rather a new philosophy of dealing with what has been regarded as wastewater
in the past. The systems of this approach are based on the systematic implementation of a
material-flow-oriented recycling process as a holistic alternative to conventional solutions.
Ideally, ecological sanitation systems enable the complete recovery of all nutrients from fae-
ces, urine and greywater to the benefit of agriculture, and the minimization of water pollution,
while at the same time ensuring that water is used economically and is reused to the greatest
possible extent, particularly for irrigation purposes”.

2.4 On-site and waterless sanitation: a component of ecological sanitation


On-site options include composting toilets and constructed ecosystems for graywater that use
up the effluent on site. In a Canadian office building and in a Swedish apartment building,
waterless composting system process toilet wastes, while graywater is filtered and used to
irrigate the landscapes around the buildings. In homes with on-site waterless toilet systems
(WTS) in Toronto, Canada and in Massachusetts, graywater is utilized by water-loving plants
in planter beds and greenhouses, which are integrated into the homes.
The United States Environmental Protection Agency has found that decentralized waste
treatment systems are a viable alternative to centralized systems when analyzing their
effectiveness and economics.
WTS (also known as dry, composting and biological toilets and non-liquid saturated sys-
tems) are gaining popularity because, among wastewater treatment technologies, they are one
292 D. Del Porto

of the most direct ways to avoid pollution and conserve water and resources. Most users who
install WTS do so simply because they need to have a toilet system where a conventional
water-borne system is unavailable or cannot be installed due to environmental constraints.
Long used by developing countries, parks, subsistence homeowners, and vacation cottage
owners around the world, WTS are now making their way into mainstream year-round homes
in North America, for many reasons:

– Micro-flush (one pint or less) toilets are increasingly used with WTS, making these systems
more socially acceptable.
– Graywater (wash water) systems are increasingly permitted by public health officials.
– Service contracts are available for maintaining WTSs.
– Water shortages threaten at least one-third of the world. Some estimates place it at one-half.
– Many states are tightening on-site wastewater system standards, so that many of the USA’s
millions of septic systems are now considered inadequate, and therefore in noncompliance.
As a result, many property owners are seeking ways to supplement their septic systems,
so they can avoid installing new ones. Diverting excrement and flush water from the flow
removes more than 90% of the pollution, leaving only graywater to manage.
– Population densities are increasing in cities and coastal areas, intensifying the challenge of
managing human waste.
– Owners are converting vacation homes into year-round residences. These homes are often in
remote and environmentally sensitive natural areas, such as seacoasts, lakes and mountains,
with limited capacity for wastewater disposal.
– Individuals and institutions are increasingly interested in sustainable technologies, as the
public’s awareness of sustainability issues grows.

2.4.1 Sewer-less society


According to the United States Environmental Protection Agency (USEPA), and the United
States Census Bureau, on-site systems are increasingly chosen over central sewer systems
by property owners and municipalities because they cost less than a central sewer system
(USEPA, 1997).
Public health specialists at development agencies worldwide are promoting effective and
ecological on-site waste treatment systems that save water and help prevent the spread of
fecal-oral disease.
At the same time, the acceptance of WTS as a technology has grown tremendously. They are
far more efficient, refined and proven. Every year, more states change laws and regulations to
permit them. Even researchers at Harvard University have decided that this is the technology
of the future, and have developed a high-tech prototype smart WTS with solid-state sensors
and microchips that control the process.
Thanks to these developments, WTS – long considered appropriate only for remote appli-
cations – may soon be widely viewed as a conventional wastewater treatment technology with
obvious advantages for the present and the future (Del Porto & Steinfeld, 1999).

2.5 Innovative financing and management


The United States Environmental Protection Agency and regulators worldwide are recom-
mending the formation of on-site management districts in response to poorly maintained or
inadequate conventional on-site systems. These would involve a central organization that
Urban and industrial watersheds and ecological sanitation 293

manages a district’s on-site systems, so no matter what system a property owner used, an
agency would be accountable for its performance.
The formation of these districts would allow on-site systems to receive the federal funds
for design, construction and maintenance that were once provided only for central wastewater
treatment plants.
Recently New York City instituted a financial incentive for individual property owners to
recycle wastewater and stormwater (New York City Department of Environmental Protec-
tion, Web Page). The program, called the Comprehensive Water Reuse Program (CWRP),
provides a 25% reduction in water rates and a reduced sewer bill to property owners who
install wastewater-recycling equipment (for onsite non-potable reuse) and rainwater collection
equipment as well as take other water-saving measures.

2.6 Comparing two water management scenarios (Steinfeld, 2004)


2.6.1 A hypothetical big pipe water recycling scenario
An urban neighborhood in New York City comprises industrial, residential, and commercial
buildings. The water supply transferred from a far distant watershed basin, is now completely
recycled water from one of the city’s treatment plants. Miles of pipes and pumping stations
carry away the wastewater to a treatment facility converted to a full recycle facility three
(and far more) miles away (1 mile = 1609 m). Stormwater is piped to rivers and out to sea. To
supply recycled wastewater back to this block, pipes and pumping stations have to be dug and
installed, with great disruption to the city, to pipe water via the existing supply lines or new
ones. Cost of the ultra-filtration and distribution is considerable.
Compare and contract the former to the following:

2.6.2 Hypothetical integrated source – and on site – water scenario


The same urban neighborhood is offered incentives to conserve and recycle water. Several
buildings are retrofitted with low-flush dual-flush urine-diverting toilets flushed with treated
graywater collected from sinks and showers. The flushed urine – which accounts for most of
the nitrogen in wastewater – is used to fertilize and irrigate surrounding landscaping, green
strips, parks, and green corridors via a hidden subsurface piping system. Excess nutrient-rich
urine is piped to greenhouses on industrial roof tops and brownfields to grow food or oils
used as a substitute for imported petroleum. Stormwater is collected from roofs, streets and
parking lots and treated and disinfected to be used for process water for flushing toilets and
urinals, evaporative air conditioning, boiler feed stocks, etc.
Graywater and combined wastewater is treated on site with a polyculture of treatment
systems, including planted systems such as Solar Aquatics/Living Machines and constructed
ecosystems (a category that includes planted evapotranspiration systems and constructed wet-
lands) that double as landscaping, recreation area, gardens, bird habitat, orchards, and public
art. These systems create an incentive for planners to create green spaces, such as terraces and
public gardens. The beauty of these ecological designs increase land values and foot traffic in
the neighborhoods.
Toilets and kitchens effluents drain to treatment tanks for these appliances only. Septage
is taken to a septage composting facility. Recipients of this composted septage (sludge) are
relieved to know that no industrial chemicals, toxics or heavy metals are in this material, unlike
today’s composted sewer sludge. Or, dry or micro-flush toilets drain to aerobic composters
which are periodically removed to a central processing facility.
294 D. Del Porto

Industrial flows with toxics and heavy metals are treated separately by systems designed
for their specific constituents (because the solution to pollution is not dilution). Cooling water
and condensate is recovered and recycled.
To minimize combined sewer overflows, the rooftops are retrofitted with green eco-roofs
to absorb stormwater and slow-release it for evaporation. Some of this water can be drained to
cisterns. This also reduces heat islands in the city, insulates buildings, and extends the life of
roofing. Raingardens, pervious paving, and other lower-impact development (LID) techniques
are used to infiltrate stormwater, reducing flows to the wastewater treatment plant (see Low
Impact Development Center, Web Page). Brownfields are used to infiltrate effluents and
stormwater. Unused buildings may house cisterns and treatment works. Marginal buildings
may be dismantled to make way for collection, treatment, and planted treatment systems.
Some uses are seasonal. Rainwater may be used during wet months and treated graywater
used during dry months.
A city management team periodically checks on the systems. This staff formerly worked
in the central plant or is funded with avoided central treatment costs.
The result: much-reduced flow to the wastewater treatment plant and reduced demand for
water. Also, this scenario embeds incentives for building users to reduce toxics and perhaps
water usage. This also makes for a more secure water system, as water supply and treatment
is no longer entirely centralized, so that disease outbreak, terrorist action, or treatment plant
failure do not affect users as much.
A decentralized approach may also offer a solution to one of the hurdles to implementing
full water recycling:
– Security against terrorist attacks on central systems.
– Liability is dispersed, so that any disease outbreak can be more isolated.
– Users of these water sources, now more aware of and invested in their sources, are far more
confident in using recycled water and more likely to safeguard it.
It is clear that combining excreta, toxics, heavy metals, and many other constituents in
a water-carriage wastewater infrastructure and then attempting remove it all with advanced
filtration may be unfeasible for many developing countries and perhaps the first world too.
The advantages of an integrated site- and source-based approach are too great to ignore.

2.6.3 Costs for an integrated urban water management plan


A typical Los Angeles single-family household pays only US$ 63 per month for combined
water, stormwater and wastewater services. It is estimated to cost an additional US$ 45 per
month to implement an integrated water management plan (City of Los Angeles, Web Page).

3 CONCLUSION

Many of the water crises in urban areas are based not so much on the availability of water
but rather on our failure to manage how water is used. Specifically, the failure to recognize
stormwater as a resource and the practice of using water to transport unwanted wastes.
Wastewater recycling will be a necessary strategy in the years ahead. However, it is only
one component that should be part of an integrated strategy that makes best use of local sites,
as well as local sources of water and local effluents that can be used.
Urban and industrial watersheds and ecological sanitation 295

Continuing to combine wastes – including excreta, toxics, and heavy metals – in wastewater
flows, then using advanced ultra-filtration at the end of the pipe to treat them will not be feasible
for many cities, especially those in developing countries.
To identify wastewater recycling as the best solution to the world’s water supply challenges
is to use the same mentality that created today’s many water problems.
The common urban approach of centrally collecting combined effluents from a wide variety
of sources, then treating this soup with end-of-pipe solutions, has evolved in an effort to avoid
complexity. Yet nature’s model shows us that complexity is the best way to manage resources.

REFERENCES AND WEB PAGES

Brooks, D.B. (2003). Another path not taken: a methodological exploration of water soft paths for Canada
and elsewhere. Report to Environment Canada. Friends of the Earth Canada, Ottawa, Canada.
California Department of Health Services (Web Page). Drinking Water Action Levels. Drinking Water
Program, Division of Drinking Water and Environmental Management. [http://www.dhs.ca.gov/ps/
ddwem/chemicals/AL/actionlevels.htm]
City of Los Angeles (Web Page). Integrated Resources Program. Information sheet of the Inte-
grated Resources Plan (IRP). Bureau of Sanitation Department of Public Works Program.
[http://www.ci.la.ca.us/SAN/irp/documents/factsheet.pdf]
Del Porto, D. & Steinfeld, C. (1999). The Composting Toilet System Book. The Center for Ecological
Pollution Prevention, Concord, Massachusetts, USA: 235 pp.
Del Porto, D. & The Ecological Engineering Group (2001). Consulting report for Ford Motor Company.
Concord, Massachusetts, USA.
Detroit News Online (Web Page). Detroit Water Department in midst of $4.3 billion construction effort.
2 May 2001. [http://detnews.com/2001/detroit/0105/04/s06-219086.htm]
Environment Canada (Web Page). Industrial use. The management of water, Freshwater Website.
[http://www.ec.gc.ca/water/en/manage/use/e_manuf.htm]
GTZ (Deutsche Gesellschaft für Technische Zusammenarbeit) (Web Page). Ecosan: ecological
sanitation. [http://www2.gtz.de/ecosan/english/subject.htm]
Low Impact Development Center (Web Page). Beltsville, Maryland, USA. [http://www.lid-
stormwater.net]
New York City Department of Environmental Protection (Web Page). City Introduces Innova-
tive New Comprehensive Water Re-Use Program. DEP News, Press Release, 2 April 2004.
[http://www.nyc.gov/html/dep/html/press/04-16pr.html]
Steinfeld, C. (2004). Scenarios for Urban Integrated Water Planning.
United Nations (2004). UN Report says world urban population of 3 billion today expected to reach
5 billion by 2030. Press Release, POP/899. 24 March. Department of Economic and Social Affairs
(DESA), United Nations Population Division, New York, USA. [http://www.un.org/esa/population/
publications/wup2003/pop899_English.doc]
USEPA (United States Environmental Protection Agency) (1997). Response to Congress on Use
of Decentralized Wastewater Treatment Systems. USEPA Office of Wastewater Management,
Washington, D.C., USA. Available at: [http://www.epa.gov/owmitnet/mab/smcomm/scpub.htm]
Vickers, A. (2001). Handbook of Water Use and Conservation. Waterplow Press, Amherst, Mas-
sachusetts, USA: 464 pp.
WateReuse Association (Web Page). WateReuse Association testifies before Subcommittee on Water
and Power. 15 April 2003. [http://www.watereuse.org/Pages/pr041503.htm]
CHAPTER 18

The potential for desalination technologies in meeting the


water crisis

J. Uche, A. Valero & L. Serra


CIRCE Foundation, University of Zaragoza, Spain

ABSTRACT: This chapter presents the potential paper of desalination as a new and non-finite source
of fresh water that plays an important role in solving fresh water scarcity in some oil-rich regions and the
Middle East, and in many arid and semiarid countries. The most widespread desalination technologies,
the weight of desalination in the different regions of the world and the future trends, the economic costs
of desalted water and the uses derived from those costs, and the environmental charges of desalination
and their corrective measures to be undertaken have been reviewed carefully in the paper. Finally,
the future improvements expected for desalination technology, and their economic and environmental
consequences are also briefly commented in the chapter.

Keywords: Desalination, water resources, seawater, brackish water, brine

1 INTRODUCTION

Water scarcity will soon be a serious problem, especially considering the rapidly world popu-
lation growth and raising water consumption per capita. So, it is becoming an increasing
problem in many regions of the planet. For any society, 1000 m3 /yr per person is considered
the standard benchmark level below which chronic water scarcity is considered to impede
development and harm human health. However, we are using now about 4000 km3 /yr (World
Water Council, 2000). The picture of the present situation is really dark, as figures show that
30% of the world population is suffering water stress, 3000 million people lack an appropriate
water sanitation and 12% of the world population uses about 85% of the water supplied
(Freshwater Action Network, 2002), while the world’s renewable fresh water resources than
can be technically used are about 10,000 km3 /yr (Shiklomanov, 1998). All these data clearly
indicate that fresh water demand will increase in the next years.
Desalination, which is of continuous increasing importance, is a possible and very interest-
ing alternative for increasing fresh water resources, and it is already a means of augmenting
them in many parts of the world, especially in the Middle East but also for the European Union
and particularly in the Mediterranean countries and island territories. By the beginning of the
year 2002, a worldwide total of 15,233 desalinating units with a total capacity of 32.4 Mm3 /d
of fresh water have been installed or were under construction (Wangnick, 2002a). That repre-
sents more than 30-fold increase in global capacity over three decades. Yet desalinated water
is only about 0.2% of the fresh water used worldwide.
As Leon Awerbuch (2002) comments, “Desalination is the only realistic hope to create new
water resources in the midst of the doom and gloom of water crisis and water pollution. The
298 J. Uche, A. Valero & L. Serra

19th century was the century of gold. The 20th century was driven by oil. In the 21st century,
water will be the most important resource. Desalination will deliver the promise not only to
create new water but also to produce fresh water at dramatically reduced cost”.
Nevertheless, desalination is an energy intensive process. Producing the total amount of
desalted water in all of the facilities in the world requires a huge amount of energy, which is
approximately equivalent to the 0.3% of primary energy in terms of fossil fuels consumed all
over the world (Uche et al., 2003). To underline how important energy is in desalination, if all
the water consumed in the world came from desalination plants, the required oil would surpass
the current yearly oil consumption. The world’s overall energy demand, 80% proceeding from
fossil fuels, is continuously increasing and it will be maintained in future. Environmental
problems associated to that energy consumption should also take into account in the near
future. Thus, decreasing the production of whatever good or commodities, e.g. fresh water,
also reduces the energy consumption. The present situation demands for imaginative solutions,
for instance process integration or the use of renewable energy sources to produce desalted
water, in order to prevent from more profound economic and social effects. Unfortunately,
the development of renewable-driven desalination is still severely impeded (if not stopped)
by the pressure from contemporary economic factors and political inertia. If our technology
continues along the present unsustainable path, not only it is essential to have an orderly
transition in the energy used for desalination (from fossil fuels to renewable resources) but
the whole industry needs to gear itself towards enhanced efficiency, waste minimization and
less environmental impact.
This lecture reviews the state of the art of desalination technology, including the most
widespread technologies, the countries in which desalination is really an important portion
of their water resources, the economic trend of desalted water produced and that will be
produced in the next future, and obviously the environmental loads of producing desalted
water by different methods. Finally, the lecture will introduce the improvements that allow
desalination to be introduced in areas where nowadays it is not considered.

2 DESALINATION TECHNOLOGIES

All major desalination technologies have demonstrated their ability and reliability to produce
fresh water in an economical manner. Each technology has found their supporters and users,
stimulating competition between them and provoking the continuous improvement of all of
those technologies.

2.1 Multistage flash (MSF) distillation (Figure 1)


Multistage flash distillation is the most widely form to produce water from seawater, espe-
cially common wherever the temperature, salt content, biological activity or pollution level
of seawater is high, as in the Middle East. This process has been in large-scale commercial
use for over 30 years, coupled to power stations. In general, MSF plants are more common
because they are simple and robust, although their specific consumption may be higher than
other technologies. Other advantage of MSF plants are their unit size, considered of large
scale (more than 50,000 m3 /d of capacity per unit).
The Multistage Flash process is described as follows: seawater pumped through heat
exchanger tubes installed in the various evaporator stages, is heated to a certain temperature.
The potential for desalination technologies in meeting the water crisis 299

Figure 1. Sketch of a typical MSF unit.

Final heating is performed by steam (coming from a power station) in a brine heater. The hot
seawater then goes into flash chambers where the pressure is maintained below the equilibrium
pressure corresponding to the temperature at which the brine enters. Part of the brine flashes
into vapor and after passing a demister, it condenses outside the tubes while heating the sea-
water flowing through the tubes. The multistage flash distillation unit contains cells assembled
in series, at a different pressure. The water produced in each stage is collected in a trough
mounted below the tube bundle which collects the fresh water end product. These widely used
units perform recycle brine in order to reduce the quantity of the make-up seawater needed to
produce fresh water. The concentrated seawater is also removed from the last stage by a pump
or by gravity.

2.2 Multi-effect distillation (MED) (Figure 2)


Contrary to MSF, in Multi-effect distillation evaporation takes place on surfaces, by exchan-
ging the latent heat through the heat transfer surface between condensing vapor on one side
and evaporating brine on the other.
The MED plant also has several stages, each with a heat exchanger tube bundle. Seawater is
sprayed onto the tubes and the condensing heating steam inside the tubes evaporates part of the
seawater on the outside. The steam produced is used as heating steam in the next stage, where
it condenses inside the tubes. The condensate is the water product. Obviously, the boiling
temperatures and pressures in the different evaporators cannot be the same. The first stage is
heated by external steam from a heat recovery steam or a back-pressure steam turbine, but in
most cases, MED plants are equipped with thermal vapor compressors for better efficiency.
The steam produced in the last stage is condensed on the outside of exchanger tubes in a
separate condenser, which is cooled by incoming seawater. Part of the heated seawater is then
used as feedwater. Product water and concentrated seawater are then pumped out from the last
stage of the evaporator.
300 J. Uche, A. Valero & L. Serra

Figure 2. Scheme of a MED unit with thermocompressor (TVC).

The major advantage of MED with respect to MSF is the ability to produce more water per
steam consumed (Performance Ratio). MED plants could reach a value of 15 while MSF almost
10, and includes lower specific power consumption (<1.5 kWh/m3 ) than MSF (>3 kWh/m3 ).
Furthermore, their efficiency does not depend on the steam temperature coming from the
turbines as MSF plants, so MED plants are the most promising distillation technologies for
desalination. However, the unit size of MED plants is up to the third with respect to MSF units
at present.

2.3 Vapor compressor (VC) distillation (Figure 3)


Vapor compression distillation is similar to multi-effect distillation, but the main difference is
that the vapor produced by the evaporation of the brine inside is not condensed in a separate
condenser. In this case, that vapor enters in a centrifugal, single-stage type designed for high-
volumetric flows, and this high-energy compressed steam is discharged into the evaporator
onto the outside of the enhanced surface tubes, where it condenses and provide its latent heat
energy to the boiler seawater inside the tubes.
Note that the process is very efficient thermodynamically, because most of the shaft work
required by the compressor is used to avoid the boiling point elevation of seawater. In com-
parison with thermal desalination plants, no cooling water is required resulting in smaller
intake and pumping systems and lower energy requirements (up to 9 kWh/m3 of product).
Unfortunately, the maximum size of VC plants is only about 3000 m3 /d but they could grow
in capacity and number of effects in order to improve its efficiency and reduce its cost.

2.4 Reverse osmosis (RO) (Figure 4)


Seawater reverse osmosis plants allow to demineralize water in a reliable manner. In RO desali-
nation, seawater (or brackish water in case of inland territories that suffer from saline aquifers)
is pretreated to avoid membrane fouling. It then passes through filter cartridges (a safety
The potential for desalination technologies in meeting the water crisis 301

Figure 3. Typical one-stage vapor compressor (VC) distillator.

Figure 4. Diagram for a seawater reverse osmosis (RO) unit.


302 J. Uche, A. Valero & L. Serra

device) and is sent by a high-pressure pump through the membrane modules (permeators).
Because of the high-pressure, pure water permeates through the membranes and the seawater is
concentrated. The water product flows directly from the permeators into a storage tank, and the
concentrated seawater (at high pressure) is sent via an energy recovery system back into the sea.
Seawater RO process uses only electrical energy, and the higher consumption occurs in the
high-pressure pump. Nowadays, energy recovery systems, as those:
– Pelton or Francis turbines, which takes profit of the pressurized brine as water stored in
reservoirs.
– Inverse pumps (or turbochargers), that allow to impulse a great part of the energy required
for the high-pressure pump, therefore they are mounted in a unique shaft.
– Pressure exchangers, translating the brine pressure into the feed water, by means of ceramic
rotors or a set of valves and closed cylinders.
Could recover more than the 95% of the energy stored in the reject brine, so the total energy
consumption for seawater could be from 3 to 5 kWh/m3 , depending on the feed water salinity
and temperature. The modularity of RO systems also provokes that those systems have been
rapidly installed in all over the world.
Seawater RO is a membrane technology. Other membrane processes as nanofiltration
(NF), ultrafiltration (UF) and microfiltration (MF) are rapidly growing to be used in brackish
desalination or waste water reuse, with very low energy consumptions (less than 1 kWh/m3 ).

2.5 Electrodyalisis (ED) (Figure 5)


This process is used to demineralize brackish water by making different ions migrate through
selective membranes in electric field made by the direct difference of voltage potential between
two electrodes connected at the boundaries of the membranes. Whenever salt water is flowing in
a cell, the cations are attracted by the anode and the anions by the cathode. If not constrained,
these ions discharge on the electrodes of opposite sign. In return, if a set of selective and
permeable membranes is placed between the electrodes, salt concentration decreases in some
compartments of the cell where salt water becomes even more concentrated. This process is
suitable for desalinating brackish waters with an average salt content between 1 to 3 g/L (for
other salinities the process is not profitable) with a very low power consumption (less than
1 kWh/m3 ) and a salt rejection of more than 80%. For seawaters, the process is also feasible
but with very high energy costs, so it is generally discarded for those proposes.

2.6 A comparison between desalination processes


Each desalination process is highly recommended in one area, but under some other circum-
stances should also be convenient to replace the previous process with some others. The Table 1
resumes the main characteristics of the referred processes.

3 DESALINATION IN THE WORLD AND FUTURE TRENDS

The information about the state of the art of the world market for all types of desalination
plants can be found in the IDA Worldwide Desalting Plants Inventory, compiled by Klaus
Wangnick (2002a). The last inventory report (no. 17) shows that by the beginning of the year
The potential for desalination technologies in meeting the water crisis 303

Figure 5. Principle of operation of a single electrodyalisis (ED) cell.

Table 1. A comparison between the most widespread desalination processes.

MSF MED VC RO ED

Energy required thermal thermal mechanical mechanical electrical


Operation temp. (◦ C) 110 70 70 45 45
Power cons. (kWh/m3 ) 3–5 1–2 8–12 3–6 0.8–1.5
Raw water quality (ppm) >50,000 >50,000 >50,000 <50,000 <3000
Product quality (ppm) <50 <50 <50 <500 <500
Unit capacity (m3 /d) 50,000 20,000 5000 200,000 10,000
Plant reliability high medium low high high
Increase capacity difficult difficult difficult easy easy
Surface required high medium low low low

2002, a worldwide total of 15,233 desalinating units with a total capacity of 32.4 Mm3 /d – or
7127 MIGD (million imperial gallons per day) – have been installed or are under construction.
Seawater plants (SWDP) produced 19.1 Mm3 /d of that capacity and 13.3 Mm3 /d are other types
of water. The total installed capacity by the end of the year 2001 was 31.4 Mm3 /d (Wangnick,
304 J. Uche, A. Valero & L. Serra

Figure 6. Total worldwide cumulative contracted desalination capacity (all waters) (Wangnick, 2001).

2000) in over 14,060 units if we consider the projects already in bidding and planning stages.
The evolution of the annually contracted capacity can be shown in Figure 6.
According to that inventory, the proportions of the various processes have been changed
along the time. Now, 60% were membrane plants and the rest were thermal plants, but ten
years ago that proportion was 50% to 50%. If only seawater desalting plants are considered,
the figures were 60% for thermal plants and 40% for membrane plants, but the ratio was 80%
for thermal plants and 20% for membrane plants ten years ago. This is due to the introduction
of RO membranes valid for the Gulf seawaters. In terms of desalination processes, 42.87% of
the total installed or contracted capacity is based on the MSF process, reflecting a continuing
decline. On the other hand, RO process increased its share to almost 39%. For plants of more
than 4000 m3 /d the MSF technology raises that proportion to 57.5% and RO decreases to
27.6%. Figure 7 shows the proportion of different processes in continents.

3.1 Middle East


In the Middle East, Saudi Arabia, the United Arab Emirates and Kuwait have a dominating
role in constructing seawater desalination plants. Particularly, the Gulf Cooperation Council
States (GCC) are the biggest users of desalination technology with more than 50% of the
total capacity in the world. Thermal processes are preferred here whereas reverse osmosis
dominates elsewhere in the world. That use of thermal plants is based on the existing good
experience in the region with such processes, and the difficulty of the Gulf seawaters (high
temperature and salinity and high amount of biological matter) for using the RO process, as
well as the low cost of energy.
The Kingdom of Saudi Arabia, supported by several seawater multistage flash plants (MSF)
desalination units combined with power generation in dual purpose power and desalination
The potential for desalination technologies in meeting the water crisis 305

Figure 7. Quote of desalination processes in each continent (Wangnick, 2001).

plants, had in 2001 the 17.3% of the total installed capacity in the world. The largest plant in
the world is Al Jubail Phase II complex, with a production of 910,000 m3 /d (200 MIGD in 40
MSF units) of distilled water and 1300 MW of electric power operating since 1982.
The third producer of desalted seawater in the world are the United Arab Emirates (UAE),
with a rank of 16.3% of the total installed capacity. This figure indicates that the desalted
seawater per capita is almost 2 m3 /d in that country. The Al-Taweelah complex contains the
largest MSF units constructed up to now, with a capacity of 57,600 m3 /d (12.6 MIGD).
Kuwait is one of the GCC countries that only incorporates MSF distillers to provide fresh
water. The number of MSF units installed in 2003 was 88, with a total capacity of 1.43 Mm3 /d
(315.6 MIGD) (Al-Fraij et al., 2003) representing almost the 5% of the total worldwide
installed capacity at that year.
Some other Middle East countries as Qatar, Bahrain (with a total capacity of 335,000 m3 /d;
Khalaf, 2003) and Oman complete the set of countries that depend fundamentally of desalted
seawater resources. The majority of installations are composed by dual-purpose power and
desalination plants that allow the combined production of water and energy in the same plant.
In those cases, some vapor coming from steam turbines or heat recovery steam generators (in
case of gas turbines or combined cycles) is used to feed the MSF or MED units coupled with
the power plants. Therefore, they are considered as huge cogeneration plants providing power
and water to the end consumers. Table 2 resumes the contribution of desalination for the most
important GCC countries.
The future trend of desalination industry is marked with the privatization of the water
and energy suppliers, taking into account the continuous growing demands of water and
energy and the constraints that suffer the public financial resources. One of the most inter-
esting innovations which are being included in that area is the concept of hybrid plants
(power station + MSF/MED + RO/VCD), which combines the advantage of distillation and
membrane processes. This combination is unique in those countries because the power demand
decreases up to 40% in winter, and that power surplus could be consumed in RO or VCD units
306 J. Uche, A. Valero & L. Serra

Table 2. Water resources in the GCC countries (Al-Weshah, 2002).

Conventional resources (Mm3 ) Non-conventional resources (Mm3 )

Agricultural
Country Surface water Groundwater Desalination drainage reuse Total

UAE 185 130 405 108 828


Bahrain 0.2 100 75 17.5 192.7
Saudi Arabia 2230 3850 795 131 7006
Oman 918 550 47.3 21.5 1536.8
Qatar 1.4 50 131 33 215.4
Kuwait 0.1 160 338 30 578.1
Yemen 3500 1400 9 52 4961
Total 6834.7 6240 1850.3 392 15,318

and that additional desalted water can be stored or consumed in the summer period. The use
of softening membranes like nanofiltration (NF) or ultrafiltration (UF) is another technique
to improve the productivity of existing MSF units avoiding the scale formation at higher
operation temperatures (Top Brine Temperatures) than conventional plants.

3.2 USA, Latin America and Caribbean


IDA Inventory Report no. 16 (Wangnick, 2000) reports that USA has the second rank in the
world capacity, with 16.7% of the overall score and more than 3000 plants. However, seawater
desalination is not representative in that country whereas desalination of brackish aquifers or
degraded aquifers by seawater intrusion (ED, RO), as well as membrane softening techniques
(membrane softening MS, NF or UF) which remove bacteria, viruses and ions are also spread
in the most arid and southern coastal states, that will account for more than 45% of the nation’s
total population growth between now and 2025.
The future of desalination in the USA is very limited for evaporation techniques due to
its lack of competence with respect to membrane technologies (Birkett, 2002). Moreover, the
future of the ED/EDR is similar to the previous processes taking into account the economy of
the RO/NF/UF and MF and its advantage in stopping pathogens that ED techniques are not
capable to do. RO techniques will be widely used in big plants (more than 10,000 m3 /d) to
reduce salinity of brackish waters or reuse waste water directly. Seawater RO plants are planned
only for big plants in Texas, California and Florida. Finally, the potential of the MF/UF/NF is
huge providing high-quality water coming from degraded surface or well waters, considering
very big plants of more than 100,000 m3 /d.
Regarding the region of Latin America and Caribbean (Andrews & Verbeek, 2002), only in
the Caribbean desalination is really important to supply water driven by the growing tourism
and decreasing costs. The total installed capacity in the Caribbean is 724,000 m3 /d by the year
2000. The most important facilities are settled at the Netherlands Antilles with a capacity of
290,000 m3 /d (it means almost 1350 L/d per person), Virgin Islands (118,000 m3 /d supposing
1350 L/d per person) and Trinidad, with more than 110,000 m3 /d. In consonance with the
Canary Islands, the Caribbean has been the cradle to prove the technological developments
of desalination along the time: the first seawater RO market, the first low-temperature MED
units, the implementation of the BOOT (Build, Own, Operate and Transferred) contracts.
The potential for desalination technologies in meeting the water crisis 307

The future in this area is the continuous growth of desalination, the majority of them based
on BOOT contracts till reaching self-sufficient islands, favored by the decreasing trends of
seawater desalinated costs. A proof is the increase of the total capacity in 57% during the
period 1996–2000. That expansion provided municipal drinking water (most of them ground
waters affected by saline intrusion) as well as industrial process water in some cases (Trinidad
island).
On the other hand, desalination in Latin America does not represent an equal weight that
in the Caribbean (530,000 m3 /d), although the capacity increase in the recent years is also
significant. The two main contributions come from Chile (131,000 m3 /d, mainly for industrial
purposes) and Mexico (a capacity of 285,000 m3 /d for industrial uses and tourism resorts).
The future of desalination in that area is focused on the use of membrane systems for removing
contamination (one of the main problems of Latin America is providing a safe supply of water),
more than reduce the salt content. Seawater RO plants will growth but not significantly for
supplying industries and covering the tourist demands in the new complex appearing in that
semi-continent.

3.3 East and Southeast Asia


The majority of desalination plants in China have industrial or power generation uses (Goto,
2002) consuming brackish, river, pure or waste waters in RO processes, and only two
municipalities and two tourism facilities exist. The total capacity at the end of 1999 was
375,000 m3 /d, but only a continuous growth is foreseen in the industrial development of the
country.
Japan leads the installed capacity in the region with more than 1 Mm3 /d and more than
1200 units. Only two big plants (40,000 m3 /d and 50,000 m3 /d) are used for domestic use, but
several small plants are sometimes used in remote areas such as small islands. On the other
hand, electronics industry consumes ultra pure water and softening techniques are applied
from pure or river waters in a great scale.
Korea had 263 desalination units in 1999 and a capacity of 448,000 m3 /d without domestic
use and RO processes consuming brackish or river waters. The industry of Taiwan required,
in 1999, 104 desalination plants with a total capacity of 164,000 m3 /d, but 4 new desalination
plants are projected for industrial parks with a total capacity of 138,000 m3 /d (Hsu, 2001).
In the Southeast, only Indonesia has a great capacity (179,000 m3 /d at the end of 1999)
which is used for industrial and power applications with only one municipality.
Summarizing, the future of desalination industry in that area is focused on the industry
requiring ultra pure waters, although demand for desalination will occur in some areas of
highly dense population and high income. A great expansion of seawater desalination is
not foreseen for the next future, especially for China despite of its economic growth: huge
inter-basins transfers are projected to provide water to the northern and highly populated
areas.
The case of India differs from the other Asian countries in the sense that there are more
than 200,000 villages with inadequate drinking water, out of which about 50,000 suffer from
brackishness problems affecting a population of about 80 million. Approximately one third of
these villages are acutely affected by salinity levels above 4000 ppm. Provision of safe water
to the villages has been given high priority in recent years, with hundreds of small RO and
ED plants (10–30 m3 /d) installed in the affected villages.
308 J. Uche, A. Valero & L. Serra

Figure 8. Role of desalination in Europe (Wangnick, 2001).

3.4 Europe and Mediterranean area


As it is expected, desalination in Europe (Figure 8) is focused in the Mediterranean coun-
tries, where the climatologic conditions and the economic activities (tourism and agriculture)
demand high water volumes. Spain is the dominating country in the scene of desalination,
with almost a capacity of 1.2 Mm3 /d at the end of 2001 (Wangnick, 2002b). It is a very
mature technology that represents the 8.3% of the urban consumption, mainly in the Canary
Islands. One thing that does not occur in other places is the use of desalination for agricultural
purposes; in some cases even seawater desalination plants have been constructed to irrigate
high-profitable crops in the Southeast of Spain and the Canary Islands. The RO technology is
dominant because the two old MSF existing units have been replaced, and several innovations
have been proved, especially in the Canary Islands, where the eastern islands (Lanzarote,
Fuerteventura, Gran Canaria) depend strongly on the desalted seawater. The future in Spain
is really promising, for instance the Spanish National Hydrological Plan approved by law in
2001 includes a list of desalination plants in the Levante that adds up to 336 Mm3 /yr (MIMAM,
2003), and some other new plants in Tenerife and the Balearic Islands. And more recently, the
new Spanish National Plan, approved in June 2004, includes new 621 Mm3 /yr with 20 new
desalination facilities located in the East and Southeast of the Spanish Peninsula.
By far, the second on that list is Italy, with a total capacity of 400,000 m3 /d, most of their
plants located at the south of the country and the island territories. The third and fourth places
are occupied by two Mediterranean islands: Malta and Cyprus. In this case desalted water
is a big percentage of the total water available in those regions. Greece, Turkey, Jordan and
Lebanon are other countries in that areas that also have small RO desalination plants. However,
The potential for desalination technologies in meeting the water crisis 309

Table 3. Water supply and demand data for the North Africa states in 2000 (Abufayed et al., 2002).

Water resources (Mm3 /yr)


Water demand
Country Surface Ground Desalination Total (Mm3 /yr)

Algeria 13,500 3700 200 17,400 6100


Egypt 65,500 7400 100 73,000 70,500
Libya 300 3400 800 4500 5300
Morocco 19,500 5000 100 24,600 13,100
Tunisia 2700 1800 300 4800 2500
Total North Africa 101,500 21,300 1500 124,300 97,500

Israel has planned a major program of seawater desalination (combined with other elements)
for the years 2002–2010 to provide new 400 Mm3 /yr of fresh water from series of 6 large
plants (from a range in size 45–100 Mm3 /yr) along the coast of the Mediterranean Sea, to
partly balance the water scarcity problems in the country (Shamir, 2003).
Russia only has a reduced capacity of their desalination plants with a total value of about
100,000 m3 /d composed by several old MED plants located at the Caspian and Black Sea. In
Central Europe, Germany, Austria and Netherlands have several plants to recycle wastewater or
produce pure water for industrial processes including power generation. They do not produce
drinking water.
Water resources in the North Africa region seem to be limited in time and space, unequally
distributed, and remote with respect to centers, suffering from a continuous increase in demand.
The geographical situation could be characterized as similar to the Middle East, but seawater
water production there is almost negligible with respect to that area, mainly due to the limitation
of their financial constraints.
The estimated cumulative installed desalination capacity reached 1 Mm3 /d at the end of
1995 (Abufayed et al., 2002). More than 625,000 m3 /d are located in Lybia (mostly MSF
plants for municipalities), and 192,000 m3 /d in Algeria mainly for industrial purposes. Egypt
incorporated 95,000 m3 /d at the end of 1999, with small RO plants and MSF/VCD plants, and
finally Morocco and Tunisia reported respectively 64,000 m3 /d and 47,000 m3 /d of installed
capacity at the end of 1995. In this regard, five North African countries (Morocco, Algeria,
Tunisia, Lybia and Egypt) requested in 1989 technical assistance from the InternationalAgency
of Atomic Energy (IAAE) to study the feasibility of desalination using nuclear power. The
future of desalination in this area is also focused on the tourism locations (Tunisia and Egypt)
that even include brackish water plants, and the increasing applications of the solar energy in
remote and dessert areas for small locations. To sum up, desalination should be the alternative
saving solution when the mobilization of non-conventional water resources is impossible or
very costly (essentially in coastal zones). Table 3 presents relevant data for the North Africa
region corresponding to the year 2000.

4 DESALINATION COSTS

The economics of desalination vary from country to country, depending upon the opportunity
cost of energy and capital as the main factors of production, and the type of desalination
310 J. Uche, A. Valero & L. Serra

Table 4. Investment of seawater desalination plants depending on


the type of process and expected plant size (Uche et al., 2002).

Specific investment
Technology Capacity range (m3 /d) cost (a/m3 /d)

MSF 10,000–50,000 1680–1080


MED-TVC 5000–20,000 1080–800
VC 1000–5000 1500–1020
RO 10,000–100,000 900–550

process. In case of dual-purpose plants, in which water and power are two separate products,
there are several methodologies that allow quantifying the costs of water and power. Moreover,
the progressive privatization of the public sectors in the Gulf Cooperation countries (GCC) had
provoked the injection of multi-billion investments coming from the private sector, allowing
under different contacts the operation and control of new big power and desalination plants
or rehabilitated existing ones. In this way, the new independent water and power producers
permits the fulfillment of the continuous growing demand for water in countries with financial
constraints.

4.1 Investment costs


Breaking the desalination costs of water, the total costs can be divided in capital costs (up
to 40% in worst cases) for interest and depreciation due to the plant investment, and the rest
are considered running (or operating) costs. The investment costs for seawater desalination
plants have decreased over the years thanks to the still ongoing development of processes but
also components and materials, apart from the increasing competence between desalination
companies. Specific costs close to 1000 a/m3 /d could be obtained for large plants in the Middle
East, whereas the same plant for RO process in some other places like the Mediterranean
countries could be constructed below 600 a/m3 /d, if no additional requirements are demanded.
Table 4 shows the expected range for different desalination processes and plant sizes.
The investment costs of the plant can be broken into the following sub-systems:
– The raw water supply system includes the seawater intake structure. It could suppose a high
percentage if the specific seawater flow is high and the sea shore is very shallow (up to 30%),
but they could be reduced if the water intake is shared with other plants, as for instance
the cooling system of the required power plant or hybrid plant. In case of hybrid plants,
the feed water for the RO plant could be the cooling water rejected from the power plant
(it has been demonstrated that RO units consume less energy when seawater is preheated,
or alternatively more permeate is produced consuming the same energy), and therefore the
cost is reduced significantly.
– The pre-treatment system is very low for thermal plants and consists of simple equipment to
storage and dose some additives (antiscale, antifoam and sodium bisulphite). However, in
RO plants costs are considerably higher because the raw seawater must be carefully treated
(chlorination, filtration, acidification, inhibition by polyphosphates and dechlorination)
before entering the RO racks. Those costs could even increase if membrane techniques are
included to reduce some pollutants in raw water. In that case it could reach up to the 20%
of the investment costs.
The potential for desalination technologies in meeting the water crisis 311

– The desalination system, by far the most expensive system of the plant, differs considerably
with the type of process and plant size. For thermal plants the elevated heat exchange area
and fine materials required implies that those costs contribute to more than the 70% of
the investment. The weight of the membrane costs and the high-pressure pump, as soon as
the energy recovery system for RO plants, do not sum up the same contribution that the
observed in thermal plants (a maximum of 60%).
– The brine disposal system includes not only the brine discharge for thermal plants, because
it exists the cooling water outfall (with identical salt content that raw water) and therefore
it is not representative for that plants. However, for RO plants, brine discharge is almost
double concentrated than raw water; specific design is required. In case of endemic species
nearby the cost, as for instance in the Spanish Levante (Posidonia Oceanica), the correction
measurements needed to protect the local ecosystems could increase the cost of that system
up to 5–7% of the total investment.
– The post-treatment system applied to the produced water is quite simple but it is obliged
by the purity of desalted seawaters. Usually, product water is re-hardened by adding carbon
dioxide and lime, and finally is disinfected with chlorine. In some cases, caustic soda is
also added to increase pH, so compensating the effect of the acid previously dosed. In both
cases (thermal or RO plants), this contribution is not significant to the total investment cost
of a desalination plant.
– The storage system of the drinking water produced consists of cylindrical tanks with a
capacity up to 24–48 hours of water production. Those tanks could contribute less than the
3% of the total investment required.
– The auxiliary equipment (instrumentation, electrical connections, air and oil systems) as
soon as the civil works required for a desalination plant does not influence the final costs
of the plants in both cases (thermal and RO plants). It could suppose up to the 10% of the
total investment of the plant.

4.2 Operating costs


The long list of components that contains the operating (or running) costs are:
– Energy cost is usually the dominant component (more than 50% of the total operating costs),
although its weight depends on the location of the plant. In oil-rich countries, energy prices
are much lower than in other areas, but in countries where energy production is expensive,
RO is always the best solution (as for example in Spain). Note that energy prices could be
increased in the near future. One of the main reasons could be the penalties imposed to
power producers in order to fulfill the Kyoto Protocol (limited CO2 emissions per country
with respect to its production in 1990).
– Personnel costs decrease as the plant size grows, except for very small plants in where
no personnel is dedicated to maintain the plant operation or shared with other activities.
Large plants are generally controlled by computer systems, and therefore people working
in those plants should be highly qualified, in order to properly manage and protect those
large investments.
– Chemical costs do not suppose an important item for thermal plants, since only antiscales
and antifoams are added to raw water. However, for membrane plants that item increases
because of its weakness to hard waters, despite of using MF or UF techniques could reduce
that operation cost but increase the corresponding investment. In the same manner, the
312 J. Uche, A. Valero & L. Serra

costs of chemical cleanings are higher in case of membrane plants, although that measure-
ment is only adopted in the worst cases, after clean washes with brine at counter-current
flow.
– Membrane replacement in RO plants, which occurs once each 3–8 years for seawater plants,
could suppose an important item if we take into account the percentage of the membranes
costs in the investment required. The rest of consumables are almost economically negligible
in membrane but also for thermal plants.
– Finally, the annual costs for spare parts can be calculated from 1% to 3% of the total
investment, depending on the plant size.

4.3 Total costs


Summarizing, the total (investment + operating) costs for a typical thermal and membrane
plant (RO) are depicted in next figures, remembering that costs could vary from country to
country.
Some examples of the total costs obtained in different locations are summarized in the
next list:

– The levelized unit cost of water in the Kingdom of Bahrain (a total desalination capacity
of 335,000 m3 /d) rises from 0.66 to 1.10 US$/m3 during the period 2003–2020, taking into
account the investments up to now and an energy cost of 0.95 US$/GJ (Khalaf, 2003).
– The first Independent Water and Power Project (IWPP) in Abu Dhabi (Al Taweelah A2)
will produce 227,000 m3 /d (50 MIGD) in a MSF plant at a cost of 0.70 US$/m3 (Awerbuch,
2002).
– The Shuweihat IWPP plant under construction, based on the BOO (Build-Own-Operate)
basis will provide 455,000 m3 /d (100 MIGD) of desalted water at a quote of 0.69 US$/m3
(Awerbuch, 2002).
– The Askhelon RO plant, with a capacity of 100 Mm3 /yr, is a BOT (Build-Operate-Transfer)
contract for 25 years at the lowest costs ever offered with that type of contracts for desalted
seawater: 0.527 US$/m3 (Lokiec & Kronenberg, 2003). Regarding the six plants included
in the National Plan 2002–2010 in Israel, water costs at the plant will range between about
0.50 to 0.55 US$/m3 at the smaller ones (Shamir, 2003). Obviously, it is not included the
cost of the conveyance system for getting water to the customers.
– Private companies have now bid less than 0.50 US$/m3 for supplying desalted water to
Singapore (Biswas & Tortajada, 2003).

Table 5 resumes some of the recent BOOT (Build-Own-Operate-Transferred) projects.


A recent list of the last large seawater desalination plants (SWDP) under biddings projected
in Spain is presented in Table 6.
It shows very clear that technological developments have consistently reduced the costs of
desalted water: in the last 20 years the cost of desalination in Spain was around 2.00–2.10 a/m3 ,
and now that cost is four times less.
If a new desalination plant is projected, a simple procedure for calculating the unit product
cost for that plant is the Desalination Economic Evaluation Program (DEEP) (Ettouney et al.,
2002), once main design data of the plant are given. It is based on a hybrid Microsoft Excel
spreadsheet and Visual Basic methodology, and is suitable for analyzing various desalination
and energy source options.
The potential for desalination technologies in meeting the water crisis 313

Table 5. Large seawater desalination plants (SWDP) projected or under construction with BOOT
contracts (Morris, 2004).

Tampa
SWDP Bay Trinidad Larnaca Dhekelia Singapore Askhelon Algeirs

Capacity (m3 /d) 95,000 135,000 40,000 40,000 136,000 274,000 200,000
Feedwater 26,000 38,000 40,000 40,000 40,000 40,000
salinity (ppm)
Contract year 2000 1996 2002 2002 2003
Years of contract 30 23 10 10 20 25 25
1st year price* 0.46 0.71 0.73 1.09 0.45 0.52 0.818

*A normalized cost of energy (0.04 cents of US$/kWh) is used.

Table 6. Recent seawater desalination plants (SWDP) in Spain. The


reflected values are the selected offers (Uche et al., 2002).

SWDP Almería Cartagena Carboneras

Capacity (m3 /d) 50,000 65,000 120,000


Depreciation cost (a/m3 ) 0.16 0.13 0.15
Fixed cost (a/m3 ) 0.10 0.09 0.04
Energy (a/m3 ) 0.19 0.18 0.18
Membrane replacement (a/m3 ) 0.03 0.02 0.02
Maintenance (a/m3 ) 0.02 0.01 0.00
Chemical additives (a/m3 ) 0.01 0.02 0.05
Total cost (a/m3 ) 0.51 0.45 0.44

4.4 Use of desalted water for agriculture


The basic use of desalination in the past and the near future is the production of drinking
water for population. Of course, the use of softening techniques will be increased in order
to be consumed in the different industrial processes. But the inclusion of desalination for
agricultural purposes, which consumes the 70% of the available water nowadays, is a very
promising option that has to be analyzed carefully. Up to now, it has been introduced in really
water scarcity areas (Kingdom of Saudi Arabia for irrigating cereals in the inland territories
of the Arabian Peninsula), but in this case the search for a policy of autonomy in agricultural
resources has been prevailed more than the economic rationality of those projects.
On the other hand, Spain is the leader in using desalination for irrigation. The total volume
destined to agriculture was estimated in 212.3 Mm3 (including brackish desalted waters) at
the end of 2003 (AEDyR, 2002). The Canary Islands and the Southeast of the country have
used since the 1980s desalted water to highly profitable crops produced in greenhouse farms
(mainly vegetables and fruits). In this sense, a very interesting study (Albiac et al., in press)
has been made to evaluate the economic alternatives to the Water Transfer proposed in the
Spanish National Hydrological Plan for the total volume destined to avoid aquifer overdraft
(419 Mm3 /yr) and guarantee the supply reliability (142 Mm3 /yr). Water management supply
and demand alternatives were examined, including banning aquifer overdraft, water pricing,
introducing water markets, and augmenting water supply with water from the Ebro River or
from seawater desalination. The results show that the best alternative for this paradigmatic
314 J. Uche, A. Valero & L. Serra

example is a combination of measures including water control and aquifer overdraft, water
trading between irrigation areas, and water supply expansion through seawater desalination
in the coastal counties (from the Alicante to Almería provinces). For instance, the effective
demand for desalted seawater (i.e. the water volume that produces positive net income to
farmers) in those counties would reach to 387 Mm3 , considering a definite cost (e.g. it includes
the cost of spreading water into the allotment) of desalted water of 0.52 a/m3 provided by
big constructed or projected plants in the area (Carboneras plant, with a design capacity
of 42 Mm3 /yr that could be doubled). This means that some counties as Campo de Dalías
and Campo de Níjar could absorb up to 60 Mm3 (49 Mm3 and 11 Mm3 respectively). Note
that the foreseeable costs of transferred water from the Ebro River could reach a range from
0.56 to 1.05 a/m3 in those coastal counties (CIRCE Foundation, 2003), so it is clear that
desalination is the more reasonable option in that coastal but with high-profitable crops. The
economical model developed in that area also indicates that desalination is not profitable for
inland territories and/or less-profitable crops like cereals or some fruit trees.

5 ENVIRONMENTAL ISSUES

5.1 Main impacts


The environmental charges associated to the desalination processes differ from each process
and location. Two major classifications should be made, taking into account the geography
and the dominant technology involved: the Middle East, in which distillation processes are
mainly applied in relatively closed oceans or seas and dessert climate, and the rest of the world
with RO as the dominant technology, and big oceans or seas. In both cases, two effects should
mainly be studied: the energy consumption of the desalination plant and the brine discharges
of plant, as soon as some other minor consequences derived from the chemical treatments and
cleaning systems, and the noise provoked in the installations if they are close to population.
In the Middle East zone, and particularly the Arabian Gulf, the long list or large distillation
plants (MSF) is provoking a serious environmental impact, in that area where the economic
growth of new formed countries is degrading its sustainable development (Al-Gobaisi, 2003).
Arabian Gulf has special characteristics that strengthen the problem:

– A reduced area of 100 km long and 300 km wide, and average water depth of 35 m, and an
average residence time of 2–5 years.
– An arid sub-tropical climate, very limited annual rainfall and an evaporation/river runoff
factor of 10 that explains the gradually increasing salinity (37,000–50,000 ppm).
– A region with political conflicts and the largest oil route in the world (20% of the total
production passes through the Strait of Hormuz).

In general, distillation plants discharges brine with a concentration of about


60,000–65,000 ppm starting from raw seawaters of 40,000–45,000 ppm. Therefore salt dilu-
tion in the ocean is not very problematic in this area, taking into account the scarce flora of the
Gulf soils. However, the continuous flow of saline waters in that closed sea would contribute
to the gradually increasing salinity already mentioned. Furthermore, CO2 and NOx emissions
derived from the elevated consumption of those plants, which include thermal but also elec-
trical consumption for pumping seawater, are very significant. Finally, thermal pollution is an
The potential for desalination technologies in meeting the water crisis 315

Table 7. Relevant airbone emissions produced by different desalination technologies


(Raluy et al., 2003).

MSF MED RO (4 kWh/m3 ) RO (3 kWh/m3 ) RO (2 kWh/m3 )

kg CO2 /m3 23.41 18.05 2.26 1.73 1.20


g dust/m3 2.04 1.02 3.10 2.37 1.56
g NOx /m3 28.30 21.43 5.10 3.92 2.74
g NMVOC*/m3 8.20 6.10 1.00 0.76 0.52
g SOx /m3 28.1 26.31 14.7 11.86 9.08

*NMVOC: Non-methane volatile organic compounds.

additional impact associated to distillation plants, because brine is usually discharged from
5◦ C to 7◦ C above the raw seawater temperature. In that case marine fauna is also affected in
a reduced sphere around the brine discharge pipe (or collector).
The situation in the other countries that incorporate desalination plants is really less dramatic
than the Arabic framework. The energy consumption of RO plants is really lower than one
consumed by distillation plants (in 5 or 6 times, see Table 7), and therefore the CO2 and NOx
emissions derived from that consumption do not represent a very important contribution. Table
7 presents the airborne emissions provoked by different desalination technologies, revealing
the gap between thermal and membrane techniques.
On the other hand, the effects on marine flora provoked by the brine disposal could be
important in some places. For instance, in the case of the Mediterranean Sea and especially
in the Spanish Levante, endemic specie called Posidonia Oceanica, a marine plant leaving
off-shore in a range of 5–30 m depth, is strongly protected by the European Union. It has
been demonstrated that it is very sensible to brine increments as some recent studies reported
(Buceta et al., 2003), and compensatory measures have to be taken (for instance enlarging the
brine discharge pipe up to the end of the existing prairies). Consequently, in most cases the
environmental load due to the installation of desalination plants can be removed by increasing
the cost of the brine disposal system.
Finally, brine disposal has a very problematic solution for brackish waters coming from
saline aquifers despite of its lower costs, since in most cases that discharge could even degrade
more and more the aquifer quality if it is not correctly designed.

5.2 Comparing with other alternatives


Anyway, the environmental impact of desalination should be compared with the impact pro-
voked by another alternative supplying the same quantity (and quality if possible) of water.
A paradigmatic example (see Serra et al., 2003, for details) could be the comparison of
the environmental assessment between the water provided by the Ebro River Water Transfer
included in the Spanish National Hydrological Plan (a maximum of 1050 Mm3 /yr) and the
same quantity of desalted water obtained by the number of desalination plants that could
supply that specified volumes. The environmental assessment tool employed is the Life
Cycle Assessment (LCA), which is one of the most powerful, recognized and internation-
ally accepted tool to examine the environmental cradle-to-grave consequences of making and
using products and services by identifying and quantifying energy and material usage and waste
discharges.
316 J. Uche, A. Valero & L. Serra

Table 8. Overall LCA scores for different energy consumptions of RO and the Ebro River Water
Transfer (pay-off period of 25 and 50 years) (Serra et al., 2003).

RO RO RO Transfer Transfer
Unit (4 kWh/m3 ) (3 kWh/m3 ) (2 kWh/m3 ) (50 years) (25 years)

Eco-indicator 99 GPts 2.62 2.04 1.46 1.86 2.20


Ecopoints 97 GPts 43,400 34,200 25,100 29,900 35,900
CML 2 baseline – 0.546 0.414 0.283 0.362 0.378

Simapro 5.0 is the software used to perform that complex LCA study, which includes the
environmental loads associated to the construction and assembly of the installations, their
operation and the final disposal once their useful life is finished. Some aspects that have not
been taken into account in the LCA of RO versus Ebro Transfer are the next:

– Biological effects derived from the brine disposal or the rupture of the river basin unit
provoking the loss of endemic species or the invasion of new invasive ones (as for instance
the zebra mussel), and the environmental impact of the water transfer in the Ebro Delta.
– The different quality of desalted and transferred water, assuming that both waters are valid
for human consumption in both cases.
– The secondary water distribution networks from the channel or permeate piping system to the
end-users, because of the lack of information about this issue in the National Hydrological
Plan.
– The uncertainty of the available volumes of the Ebro River in the next future taking into
account the historical data revealing faults in the recent years. This problem should be more
acute if it is also considered the effect of the climate change and the increasing demands in
the conceding basin.

The average energy consumption of the Ebro River water transfer was 2.5 kWh/m3 (cal-
culated from MIMAM, 2003) in the operation phase of the life cycle of the water transfer,
whereas the energy consumption of the RO was varied from 4 to 2.5 kWh/m3 , consider-
ing the decreasing trend of energy consumption in that technology. The Spanish model to
produce power is considered for that analysis (51.3% is thermal, 13.3% is hydropower and
the other 35.4% is nuclear energy). The overall scores obtained with the analysis are pre-
sented in the Table 8. Three different methods have been used. Note that those methods
provide a unique number without any physical sense but very useful to compare different
alternatives.
It means that with the present state of the art of both technologies, RO is slightly more
pollutant than the Ebro Water Transfer. The importance of energy can be better understood
when the LCA results are presented in percentage from each life cycle stage, as shown in the
Table 9.
In both options, the operation stage, in which the energy consumption is the most important
factor because the infrastructure and the plants are already built, is clearly the most important
phase. Thus, the efforts for reducing the environmental load in both options should be devoted
in the address of reducing the energy consumption and/or the usage of environmental-friendly
energy production systems. Anyway, the weight of the water transfer construction is quite
important with respect to the same phase for RO plants, a factor which is usually not considered
The potential for desalination technologies in meeting the water crisis 317

Table 9. Environmental load, for three different assessment methods, corresponding to each life cycle
phase for the RO (4 kWh/m3 ) and the Ebro River Water Transfer, considering the pay-off periods of
25 and 50 years (Serra et al., 2003).

Process MSF Unit Eco-indicator 99 Eco-points 97 CML 2 baseline

RO Assembly % 6.95 10.22 2.03


Membranes % 0.77 0.28 0.06
Operation % 92.28 89.5 97.91
Water Transfer Assembly % 31.58 33.6 8.58
(25 years) Operation % 68.41 66.4 91.42
Water transfer Assembly % 19.09 20.32 4.50
(50 years) Operation % 80.91 79.68 95.50

for other environmental assessment techniques. Note that the final disposal stage of both
alternatives does not suppose any contribution with respect to the other stages.
The results obtained here present a very optimistic future for desalination (especially RO)
with respect to conventional hydraulic projects, since:

– Technologic improvements permit a continuous reduction of energy consumption. On the


contrary, hydraulic projects are a mature technology.
– The gradual inclusion of renewable energies (see Serra et al., 2003, for different
energy production scenarios) favors RO desalination more than conventional water supply
techniques.
– By integrating desalination with energy production systems, a very important reduction of
the environmental charges of desalination could be obtained (Raluy et al., 2004).

5.3 Using renewable energies


The use of renewable energies for producing desalted water is not developed yet at a major
scale, but it could be a very interesting solution for remote and lowly populated areas, where no
other alternatives are available at affordable costs. The use of solar direct heating onto a glass
surfaces collecting evaporated seawater in a greenhouse recipient only produces up to 4 L/m2
of glazed cover; and the use of parabolic through collectors (PTC) concentrating the solar
energy to heat a thermal fluid or produce steam that it is consumed in a thermal distillation
process (MSF or MED small plant coupled to the energy storage system) can improve that
productivity up to 10 L/m2 (García et al., 1999).
Photovoltaic solar cells connected to small RO units is quite affordable if a storage system
and/or batteries are also included in the design project, in order to smooth the solar irradiation
along the day. Anyway, the cost of that solution is only recommendable for isolated locations
(see Espino et al., 2002). But the most promising renewable energy for connecting RO desalting
units is wind, considering than even wind farms would be settled off-shore and therefore
its environmental impact would be strongly diminished, and the great power that could be
generated in the coastal line provoked by a constant wind regime. The integration of that type
of energy with a pumping system that allows the storage of seawater in small dams that could
produce energy in a hydroelectric installation, and finally desalted RO in a continuous manner,
is a project that will be developed in the island of El Hierro (Canary Islands) and seems to be
one of the promising future line for desalination in a low-medium scale.
318 J. Uche, A. Valero & L. Serra

6 FUTURE INNOVATIONS IN DESALINATION

Dealing with the present technologies and their improvement, the future of desalination tech-
nology is addressed to the reduction of costs and the minimization of the environmental impacts
generated by those technologies.

6.1 Reducing costs


Several future research lines are opened in this field. The more important ones are remarked
here:
– The research oriented to the integration of water and energy systems, is essential to reduce
the costs of water and energy in dual purpose plants. Up to now, water and energy production
were managed independently, but a combined management and the appropriate tools as for
instance Thermoeconomic Analysis (Uche, 2000) permits an in-depth knowledge of both
plants and their interactions, thus discovering guidelines for the reduction of both costs. Use
of residual heats or exhausted flows rejected to the environment is highly recommended in
those plants.
– The use of hybrid systems (RO + MSF/MED units, in which RO only works in low-peaks
of power demand, usually in winter periods). The concept of aquifer storage could also be
combined in those periods, reducing therefore the cost of desalted water (costs are always
calculated per year of operation) (Awerbuch, 2002).
– Last idea suggests increasing the plant availability or utilization factor, permitting the
cost reduction. The use of experienced designers, consultants and operators, as soon as
high qualified personnel is also essential to maintain those high factors. Use of materials
resistant to corrosion is another factor improving that availability, especially for thermal
plants (Wangnick, 2001).
– In the MSF process, the inclusion of softening membranes (UF/MF) for allowing higher
Top Brine Temperatures (up to 140◦ C) without the problems of scaling and precipitations,
and therefore increasing the process productivity (e.g. higher Performance Ratio) (Ejjeh,
2001). Another way of reducing water costs in thermal techniques is augmenting the Heat
Transfer Coefficients (HTC) of evaporators/condensers, and therefore the required surface
area could be diminished. If surface area is maintained, seawater production can be increased
considerably on the other way (Wangnick, 2001). Alternatively, as MED plants work at low
temperatures (up to 70◦ C), the inclusion of aluminums or even plastic materials will reduce
considerably the investment cost of evaporators.
– For membrane processes, one of the prominent lines is the search for a more simplified
scheme in the pretreatment system, mainly due to the enormous space required for the
conventional treatment – flocculation, sedimentation and filtration, followed by diverse
additives including chlorination and dechlorination. And the principal equipment of RO
plants (membranes) is yet susceptible for major improvements (Fariñas, 2001): flow through
membranes could be increased (or alternatively the pressure required to produce the same
flow could be reduced), and/or reduce the specific costs of membranes (e.g. the cost per
unit of surface in spiral membranes). Remember that the pressure required to produce a
significant flow across the membrane is more than two times the osmotic pressure required
for a typical seawater.
The potential for desalination technologies in meeting the water crisis 319

Figure 9. Water cost obtained by seawater desalination in Spain along the time (AEDyR, 2002).

– The inclusion of the energy recovery systems for large RO plants (for instance the pressure
exchanger supplied by RO kinetic) substituting Pelton turbines could reduce up to 1 kWh/m3
the specific energy consumption, reaching a total consumption (intake pumping + RO
process + permeate pumping to storage system) of less than 3 kWh/m3 nowadays. In this
way, the advantage obtained by these techniques is yet close to its limit, having an efficiency
of about 95%.
All those measurements could reduce the costs of desalted water on the next future but not
as recent decades, as we can see in the historical trend for the water cost desalted in Spain
(Figure 9). This is mainly due to the compensation produced by the expected increasing prices
of energy, taking into account for the political situation and the environmental charges that
will be associated to energy.

6.2 Reducing the environmental impacts


To reduce the environmental impacts of desalination plants, starting from the present situation,
some guidelines are suggested here:
– First of all, it is essential to reduce the energy consumption for all technologies, and inte-
grate them with energy production systems, particularly with those driven by renewable
energies.
– Impede the corrosion of materials.
– Use sound-insulation materials to reduce the impact of noise.
– Reduce the addition of chemicals and use the less aggressive ones to the environment.
– Avoid (if possible) the use of chlorine in desalination processes by investigating new afford-
able disinfection methods, in order to prevent the formation of trihalomethanes (Wangnick,
2001).
– Reduce the thermal contamination by immersing cooling exchangers.
– Increase the investment required for enlarging the brine discharge pipes, also for inland
territories to collect saline waters from brackish plants.
320 J. Uche, A. Valero & L. Serra

Finally, it would be desirable to apply the Environmental Impact Assessment (EIA, usu-
ally required in most countries) for new facilities, and especially strategic EIA that consider
cumulative impacts form various sources in one region (Schiffler, 2004).

7 CONCLUSIONS

Desalination represents nowadays a reliable potable water supply, especially for providing safe
fresh water to big coastal cities. In some cases, desalination should be the dominant source for
balancing water demands of a region (see Bremere et al., 2001), but not for other ones: up to
now, desalination do not allow to solve the water problems in all over the world. The reasons
for that assumption are, among others, briefly explained and classified here:

7.1 The Integrated Water Resources Management (IWRM) solution


– Desalination should always be the last solution to avoid water availability problems: water
demand strategies as efficient irrigation methods, reduction of leakages in pipes, use of
water saving devices in households, introduction water markets in drought periods, and the
intensive use of reused water should be strengthened previously. A paradigmatic example
of that inadequate projection is Malta island (Schiffler, 2004): when leakage losses were
reduced, desalination plants had overcapacity.
– Desalination should be the selected water supply alternative if provokes less economic and
environmental charges to conventional alternatives for supplying water (other solutions
may involve inter-basin transfers with a huge budget and uncertainties derived from the
climatology).

7.2 Costs
– Currently, only 3% of global drinking water is supplied through desalination, but it is
concentrated in developed countries and in the Arab Gulf countries (remember that 2700
million people live on less than 2 US$/d) (Schiffler, 2004).
– Costs have fallen to affordable levels for many communities (up to 0.50 a/m3 ), but in the
future they will continue to fall but not as fast as previously, since no major breakthroughs
are envisaged in the immediate future.
– The use of desalted water for agriculture should be studied carefully. Usually, the economic
rationality of this solution is not justified, and the use of subsidies is quite extended. More-
over, in water scarce areas desalination for urban supply should not be applied to free up
water for irrigation.

7.3 Environmental impact


– Desalination has local impacts but they can be mitigated, usually at low or moderate costs,
which obviously must be charged to the users. And in most cases, they can be outweighed
by the benefits (for instance avoiding the use of alternative conventional resources harmful
with the environment).
– However, it has a serious global impact derived from its energy consumption. Remember
that the thermodynamic limit for desalting seawater is about 0.8 kWh/m3 , i.e. the minimum
The potential for desalination technologies in meeting the water crisis 321

realistic consumption for the complete plant would be estimated in 1.5 kWh/m3 , equiva-
lent to a groundwater pumped out from a well at about 400 m depth. Therefore, seawater
desalination will not be the definite solution for water scarce areas till renewable ener-
gies would be massively introduced (maybe in the next 20–30 years), taking into account
that 1200 million people lack potable water and 3000 million lack appropriate sanitation
conditions.

REFERENCES

Abufayed, A.A.; Elghuel, M.K.A. & Rashed, M. (2002). Desalination: a Viable Supplemental Source of
Water for the Arid States of North Africa. Desalination, 152: 75–82.
AEDyR (Spanish Association for Desalination and Water Reuse) (2002). Publicity brochure of
desalination in Spain. Presented at the IDA 2002 Conference. Manama, Bahrain.
Albiac, J.; Hanemann, M.; Calatrava, J. & Uche, J. (in press). Evaluating Alternatives to the Spanish
Nacional Hydrological Plan. Water Resources Development, special issue.
Al-Fraij, K.M.; Al-Adwani, A.A. & Al Romh, M.K. (2003). The Future of Seawater Desalination in
Kuwait. IDA 2003 Conference, Paradise Island, Bahamas.
Al-Gobaisi, D.M.K. (2003). Sustainability of Desalination Systems – an Essential Consideration for the
Future. Integrated power and desalination plants. EOLSS Publishers Ltd., London, UK.
Al-Weshah, R. (2002). The role of UNESCO in Sustainable Water Resources Management in the Arab
World. Desalination, 152: 1–13.
Andrews, W.T. & Verbeek, V. (2002). Regional Report on Desalination – Latin America & Caribbean.
IDA 2002 Conference. Manama, Bahrain.
Awerbuch, L. (2002). Vision for Desalination – Challenges and Opportunities. IDA 2002 Conference.
Manama, Bahrain.
Birkett, J. (2002). Desalination in North America. A Regional Review. IDA 2002 Conference. Manama,
Bahrain.
Biswas. A.K. & Tortajada, C. (2003). Assessment of Spanish National Hydrological Plan. Reports of
International Experts required by the Government of Aragón.
Bremere, I.; Kennedy, M.; Stikker, A. & Schippers, J. (2001). How Water Scarcity will Effect the Growth
in the Desalination Market in the Coming 25 years. Desalination, 138: 7–15.
Buceta, J.L.; Gacia, E.; Mas, J.; Romero, J.; Ruiz, J.; Ruiz-Mateo, A. & Sánchez, J.L. (2003). Estudio de
los incrementos de salinidad sobre la fanerógama marina Posidonia Oceánica y su ecosistema, con el
fin de prever y minimizar los impactos que pudieran causar los vertidos de agua de rechazo de plantas
desaladoras. Final report and recommendations. [Pers. comm.]
CIRCE Foundation (2003). Alegaciones al Proyecto de Transferencias aprobado en el artículo 13 de la
Ley 10/2001 y su Estudio de Impacto Ambiental. [Pers. comm.]
Ejjeh, G. (2001). Desalination, a Reliable Source of New Water. International Conference Spanish
Hydrological Plan and Sustainable Water Management. Zaragoza, Spain.
Espino, T.; Peñate, B. & Piernavieja, G. (2002). Proyecto Dessol: experiencia piloto de desalación de agua
de mar por ósmosis inversa alimentada con energía solar fotovoltaica. Una opción de abastecimiento
para zonas costeras. AEDyR 3rd Annual Congress. Málaga, Spain.
Ettouney, H.M.; El-Dessouky, H.T.; Faibish, R. & Gowin, P.J. (2002). Evaluating the Economics of
Desalination. Cep magazine (www.cepmagazine.org).
Fariñas, M. (2001). Novedades tecnológicas en la desalación por ósmosis inversa. International
Conference Spanish Hydrological Plan and Sustainable Water Management. Zaragoza, Spain.
Freshwater Action Network (2002). NGO Guide to the Water-Dome, the World Summit on Sus-
tainable Development and International Water Policy. World Summit on Sustainable Development.
Johannesburg, South Africa.
García, L.; Palmero, A.I.; & Gómez, C. (1999). Application of Direct Steam Generation into a Solar
Parabolic Trough Collector to Multieffect Distillation. Desalination, 125: 139–145.
Goto, T. (2002). Water Problems in East and South Asia. IDA 2002 Conference. Manama, Bahrain.
322 J. Uche, A. Valero & L. Serra

Hsu, S.K. (2001). Seawater Desalination as an Alternative Water Supply in Taiwan. IDA International
Conference. Singapore, March 2001, Session VII.
Khalaf, A. (2003). The Effective Unit Cost of Electricity and Water in the Kingdom of Bahrain. IDA
Conference. Paradise Island, Bahamas.
Lokiec, F. & Kronenberg, G. (2003). South Israel 100 million/yr Seawater Desalination Facility: Build,
Operate and Transfer (BOT) project. Desalination, 156: 29–37.
MIMAM (Ministerio de Medio Ambiente) (2003). Proyecto de Transferencias autorizadas en el artículo
13 de la Ley 10/2001 del Plan Hidrológico Nacional y su Evaluación de Impacto Ambiental. Spanish
Ministry of Environment. Madrid, Spain.
Morris, R. (2004). Technological Trends in Desalination and Their Impact on Costs and the Environment,
World Bank Water Week. Washington, D.C., USA.
Raluy, G.; Serra, L. & Uche, J. (2003). Life Cycle Assessment of MSF, MED and RO Desalination
Technologies. Second Conference on Sustainable Development of Energy, Water and Environment
Systems, Dubrovnik, Croatia.
Raluy, G.; Serra, L.; Uche, J. & Valero, A. (2004). Life Cycle Assessment of Desalination Technologies
integrated with Energy Production Systems. Desalination, 167: 445–458.
Schiffler, M. (2004). Desalination: Recent Trends and the Role of the World Bank. [Pers.comm.]
Serra, L.; Raluy, G.; Uche, J. & Valero, A. (2003). Environmental Impact of Water Production Technolo-
gies. Life Cycle Assessment of Ebro River Water Transfer versus the Reverse Osmosis Desalination.
Report financed by the Government of Aragón. [Spanish version available at www.trasvasebro.com].
Shamir, U. (2003). Review and Evaluation of Certain Aspects of the Spanish National Hydrological
Plan. Reports of International Experts required by the Government of Aragón.
Shiklomanov, V. (1998). World Water Resources. A New Appraisal and Assessment for the 21st century.
UNESCO, Paris, France.
Uche, J. (2000). Thermoeconomic Analysis and Optimization of a Dual-Purpose Power and Desalination
plant. Ph.D. Thesis. Department of Mechanical Engineering, University of Zaragoza, Spain.
Uche, J.; Valero, A. & Serra, L. (2002). La desalación y reutilización como recursos alternativos.
Administrative document printed by the Government of Aragón.
Uche, J.; Serra, L.; Herrero, L.A.; Valero, A.; Turégano, J.A. & Torres, C. (2003). Software for the
analysis of water and energy systems. Desalination, 156: 367–378.
Wangnick, K. (2000). 2000 IDA Worldwide Plants Inventory. Report no. 16. Wangnick Consulting.
Wangnick, K. (2001). A Global Overview of Water Desalination Technology and the Perspectives.
International Conference Spanish Hydrological Plan and Sustainable Water Management. Zaragoza,
Spain.
Wangnick, K. (2002a). 2002 IDA Worldwide Plants Inventory. Report no. 17. Wangnick Consulting.
Wangnick, K. (2002b). Regional Review for Europe – The Region Leads in Seawater Desalination. IDA
2002 Conference. Manama, Bahrain.
World Water Council (2000). World Water Vision, Making Water Everybody’s Business. Earthscan,
London, UK.
CHAPTER 19

The potential for desalination technologies in meeting


the water crisis: comments

E. Custodio
Technical University of Catalonia (UPC), Barcelona, Spain

ABSTRACT: These notes add some comments to the chapter by Uche et al. (this volume) on desalin-
izated water use. The comments do not refer to the desalinization techniques themselves. Raw water
for desalination is often coastal seawater, but also brackish and saline groundwaters, both in coastal
and inland emplacements. Treated sewage water, and also brackish water from other sources, such as
return irrigation flows, are susceptible of being desalinized. Desalinization technology and costs have
improved dramatically in the last few decades with respect to specific energy consumption, the processes
themselves, the materials, and the membranes when they are used. But in spite of a current, fairly well
known, and proven technology, information on water costs remains uncertain for people and even for
policy-makers due to diverse intervening economic parameters, and how they are modified to yield to
social pressure and political goals. The relative low cost of water storage implies that under normal
circumstances the most economic way of operating a desalinization plant is continuously at nominal
capacity. Return flows from a desalinization plant must be safely disposed of and this adds significantly
to water cost. Desalinization is an energy and capital intensive process, as are also long water transfers
and tertiary treatment of waste water. Produced water cost therefore is intrinsically expensive, more
for seawater than for brackish water. It becomes a new water resource in areas and for uses that can
support high prices and have water resources at hand to be desalinized. However, this is not a resource
for poor areas or for water intensive processes except if heavily subsidized. Under normal circumstances
desalinization is just one more alternative for water supply to be considered among others.

Keywords: Desalinization, sea water, brackish water, groundwater, environment

1 INTRODUCTION

The chapter prepared by J. Uche, A. Valero and L. Serra (this volume) on the potential for
desalination technologies in meeting the water crisis prove to be very comprehensive covering
the different aspects of desalinization and its role in water resources. The emphasis in these
comments will be on water resources and some related environmental aspects and not on the
technical aspects.
In this chapter, the term desalinization shall be used with preference to desalination. In
fact the Webster’s Ninth New Collegiate Dictionary (1983) says: “Desalination: see desalt;
Desalinization: see desalt; Desalt: to remove salt from . . . ”. It is clear that these are synonym-
ous terms. The preferential use of desalinization tries to point out that the focus is in making
the water less saline and fresh instead of taking out salts, although some processes accomplish
the latter. Some experts on this topic argue that desalination is the right term when raw water is
324 E. Custodio

sea water, and desalinization is preferable when raw water is brackish water. In the opinion of
the author this is too artificial and does not introduce any significant language improvement.
Desalinization is any process whose objective is to reduce the saline content of water that
is naturally rich in dissolved salts or has been salinized due to some natural and/or anthropic
processes.
Saline and brackish waters contain an excess of salts that prevent or make their use incon-
venient for the intended purposes (potable, urban, industrial, rural or agricultural). The
reduction of some constituents, such as inorganic carbon, Ca2+ , Mg2+ , Fe2+ , can be accom-
plished at low cost by means of conventional physico-chemical treatments. But the removal of
Cl− , Na+ and SO2− 4 , which are generally the main constituents of brackish and saline waters,
need desalinization treatments due to their high solubility and chemical stability. The reduction
of other constituents such as NO− 3 from anthropically polluted water may also be a comple-
mentary objective for desalinization. The reduction of organic contaminants and of bacteria
and viruses are a consequence of desalinization but not a main objective. Processes aimed at
these objectives, such as micro-, ultra- and nano-filtration through membranes, although they
share similar technologies with membrane desalinization, they are not properly such.
Desalinization technologies are well known and have been around since the late 1950s and
some of them now in use were developed in the 1970s (Custodio & Llamas, 1976: section 23).
During this time some major improvements have been made such as less energy consumption,
more suited materials, enhanced head transfer, and much better membranes. Sufficient experi-
ence has been also gained to allow the making of reliable projects and valid cost estimations.
Some of the distillation processes (MSF, MED, VC) and membrane processes (RO, ED) are
currently well established and developed, as commented by Uche et al. (this volume).
Other processes such as freezing and solvent extraction have not progressed beyond their
pilot plant stage, and perhaps will not be developed any further since no economic advantages
are expected.
Figure 1 presents a simple scheme of the desalinization processes with the main items and
stages that will be considered and discussed later on. It is quite self-explanatory.
It seems quite important to mention that fresh water and potable water availability is not
clearly related with water availability except in particular situations, such as arid and semiarid
areas, where the only solution is to cut down on water demand for human activities or import
water from other areas. In this case, the reuse of water may be significant.

DESALINIZATION

DESALINIZATION PRODUCT
RAW WATER CONSUMER
PLANT WATER
Sea water (fresh water)
Distillation
Brackish/saline Membranes
groundwater
Obtaining fresher water RETURN FLOW
Used water
Retaining ions and others (blowdown, rejection)

ENERGY Disposal

Figure 1. Simplified scheme of desalinization processes with the main items and stages.
The potential for desalination technologies in meeting the water crisis: comments 325

Oftentimes, available water is too saline or it has some solutes in excess or it is inorganically
or organically contaminated. This of course will reduce usable water resources and may even
create an understandable concern or the feeling of an oncoming water crisis. Nevertheless,
usable water can be produced using existing technology from any kind of water (brackish,
saline, contaminated), even though a large area for technical improvement still exists. But this
implies, however, that energy is available to do it and it can be done at an affordable price
in a way that does not further harm the environment. Jointly with waste disposal, this is a
major challenge, and a reason to not give much weight to Shiklomanov’s (1998) figure of
10,000 km3 /yr for total renewable freshwater resources, or to the threshold limits of necessary
available water per person, as pointed out by Llamas & Custodio (2003). Also, the demand
for water does not necessarily coincide with water needs (Merrett, 2004).
One major challenge is how can human populations have access to fresh and potable water
within their economic capability when a large fraction of them are poor, technologically
deprived and energy deficient. In these cases desalinization is of little use.

2 ROLE OF DESALINIZATION

A main result of the present situation is that technological progress and cost reductions have
allowed desalinization to be considered as an alternative for solving a given set of freshwater
demand problems. It becomes a real alternative when freshwater is an expensive local com-
modity due to the cost of making it available or due to market competition when it is scarce.
The concurrence is to respect other water sources that are also capital intensive and energy
consumption intensive, as occurs with long water transfers or the reclamation of sewage water,
and even deep aquifer development in extreme situations.
Real circumstances are quite variable, and normal situations may be changed into exceptions
when an area has a high water demand or is arid. These situations do not permit simple rules to
be established or detailed economic analyses to be requested in order to help in the decision-
making process. It may so happen that there is not an optimal solution, but rather a set of
possibilities where in some of them desalinization may play a role when brackish and/or
saline water is or can be made available. This includes used water and return flows from
agriculture, mining or industrial usage when too brackish for direct use.
The abundance of brackish and saline water in the ground, both in costal and inland situ-
ations, even if large parts are non-renewable reserves, opens a complementary desalinization
role for groundwater development. One primary problem that remains is what to do with the
return flow (blowdown or rejection). This will be considered later on.
As mentioned in the Introduction, will also be commented later on, and is well stressed
in the papers by Uche et al. (this volume) and Semiat (2000), desalinization is inherently
expensive, even with foreseeable technological improvements, unless energy is available at
very low cost, which is unlikely except in heavily subsidized situations.

3 DESALINIZATION GOALS

Desalinization, as a real alternative or complement to other freshwater sources, has three


different aspects.
326 E. Custodio

(a) Desalinization of sea water in coastal areas (including continental saline surface water
bodies). Raw water is either sea water directly uptaken from the sea or groundwater that
is directly recharged into the ground from the sea to avoid turbidity, solid particles and
biological problems. Usage in this situation usually takes place near the coast. Inland
transportation may imply an important added cost, especially if water demand is in high
elevation areas or far away. Every 100 m elevation or head loss adds about 0.4 kWh/m3 .
The blowdown brines (return flow) can be disposed of into the near-by sea (see later on).
(b) Desalinization of brackish and saline groundwater, either near the coast or at a distance, in
continental emplacements. Desalinization of brackish water by means of membranes has
a relative low cost and a reduced energy consumption. Its feasibility will depend on avail-
able brackish and saline water reserves or the rate at which they are produced in the case
of seawater intrusion. Feasibility will also depend on the existence of difficult-to-pretreat
inconvenient solutes, any possible side-effects of groundwater exploitation on existing
freshwater resources, and on land subsidence. The possibility of obtaining raw water and
its cost will also be another feasibility factor. Every 100 m elevation above the dynamic
groundwater level adds about 0.5 kWh/m3 to the desalinization cost for the common energy
efficiency of submerged pumps. One of the main problems in inland areas is return flow
disposal, which may vary from saline water to a brine. This may add a significant cost
increase and may even make desalination unfeasible. If there is not a deep saline aquifer
to inject them into, return flows must be carried out to the coast or to evaporation ponds
by means of brine-ducts. Brine-ducts are prone to failure and breakdown, with the added
risk of polluting continental water resources. Abstraction of saline water from a coastal or
continental aquifer may help in the protection of existing freshwater resources. This effect
must be taken into consideration in water balances, and it may affect freshwater avail-
ability if wells produce brackish water through the mixing of freshwater with more saline
water.
(c) Improving the quality of treated sewage and used water by removing a part of the salts
as well as dissolved organic matter, bacteria and viruses at the same time through the use
of appropriate membrane technologies. The cost will depend on raw water salinity and
the degree pre-treatment to be applied, and on the final quality to be obtained. Product
water may be for direct use (irrigation, industry, municipal, even household) or for further
treatment and storage by means of artificial recharge into the ground.

4 ENERGY FOR DESALINIZATION

One of the main drawbacks for desalinization is the high energy consumption and the associated
environmental problems produced. Energy production for desalination adds to the increased
impact on the Earth’s thermal balance through the release into the atmosphere of greenhouse
gases and particles. But long water transfers may also be energy consumption intensive (up to
some kWh/m3 ).
Current reverse osmosis processes with return flow pressure recovery may consume 3.0
to 4.0 kWh/m3 of product water for raw water from the sea, and proportionally less for lower
salinity raw water. These values approach the thermodynamic minimum energy for conversion.
From here on, only relatively small improvements in energy consumption can be expected for
membrane technology. There is still some more room for distillation processes. MED scores
The potential for desalination technologies in meeting the water crisis: comments 327

are also currently good for sea water desalinization in large plants, although they cannot
compete with membrane processes for brackish water.
To the cost of in-plant product water, the cost of elevation and distribution to the supply
centres, which is far from being negligible, must be added.
The price of energy will vary from region to region and may include economic charges,
taxes and subsidies of many kinds. This makes real circumstances quite variable. The future
trend is probably on the increase. In many cases the relatively cheap hydropower potential
is already fully used, fossil fuels are being depleted and renewable sources of energy are
expensive (or heavily subsidized) and represent only a relatively small contribution to total
energy demand. Variable renewable energy sources, such as wind and solar power, are poorly
suited to the capital expensive desalinization plants if used directly.
Nuclear power is a real alternative, although not a cheap one and with limited resources,
but it suffers from a wide and sometimes irrational social rejection. This has been a serious
brake for solving energy problems in many countries, and as a consequence adds to the energy-
related problems of desalinization. From the point of view of global environmental impact on
the climate, nuclear energy is potentially much cleaner than fossil fuel energy. The future for
relatively clean and affordable energy prices still relies on nuclear energy. But the gap between
the present situation and what could be a significant contribution for the future is widening.
The need is for a bold decision now and a sustained maximum effort during a series of years.
For improved nuclear fission energy production using breeder reactors, too much time has been
lost on moratoria. Any significant impact on energy availability may come too late. Nuclear
fusion reactors have a more important potential, but just to be practical this technology will
probably need more than 25 years to achieve a significant impact on energy availability.

5 DATA ON DESALINIZATION COST

When comparing desalinization with other freshwater supply sources in areas where water
prices are sufficiently high, very diverse figures are found in both technical and mass media
reports. These differences not only misinform citizens but they also mislead anyone hav-
ing any sort of decision-making capacity. Assuming that figures are not expressly modified
or falsified, which does occur sometimes to reach some pre-established result or support
some pre-established solution, the different results released for public information are quite
widespread. In addition to technical issues, the basis for economic calculations may also
clearly differ. Some of the main factors for this are the interest on money and the discount
rate to be applied, the duration of the pay-off period and the financing conditions, which
may conspicuously change from case to case. This is often due to differing public and private
investment viewpoints grounded on social interest and enterprise benefits. Public sponsoring
tends to apply lower interest and discount rates, longer pay-off periods and improved financing
conditions through hidden guarantees and diverse kinds of subsidies. Crucial also is knowing
where the water cost is calculated: at the desalination plant, or at distribution storage reservoirs,
or at the mains, or where consumed. Water quality is another factor to be considered, since
pre-treatments are expensive, a low salinity product will allow increasing water availability
by mixing with other brackish waters or waters with excessive contents of some constituents.
The manner in which taxes, fiscal exemptions, subsidies and special prices are applied also
play a significant role on the resulting costs and the economic parameters used to compare
328 E. Custodio

projects and alternatives. Publicly sponsored activities often take advantage of lower taxes,
low interest loans, subsidies on energy, land purchases or construction, increases in the price
of other commodities (e.g. electricity and motor fuels), in order to compensate for water cost
if considered too high from a social perspective. Sometimes brine-ducts are built as a common
interest investment (e.g. paying others).
These and other similar considerations mean that comparisons can only be carried out fairly
when economic circumstances are similar or respond to well-defined protocols.
A continuous source of debate about these different circumstances has taken place in the
Eastern Canary Islands, Spain, where desalinization plants of very different ages and sizes
are financed, operated and maintained from a wide combination of public and private initia-
tives. There and in other areas of Eastern Spain water prices charged to customers or used
for accounting purposes may differ by more than a factor of two. For a large reverse osmosis
plant about to be commissioned in Eastern Spain, free-market product water cost at the plant
is calculated at about 0.9 a/m3 but will be sold at 0.4 a/m3 , the difference being compen-
sated by subsidies and other economic advantages. Currently, in the midst of a controversy
between a long transfer canal and desalinization, the media is showing cost informations
that varies between 0.3 a/m3 and 1.2 a/m3 , and are seldom complemented with further data,
such as the raw water origin and salinity (brackish or marine). Only experts have some
information.
Other factors that are not easily obtainable are the applied repayment rate and time of
facilities, interest rates, the point where water is made available, cost sharing with other uses
(energy, water production), return-flow disposal and associated storage, transportation facil-
ities, energy costs and subsidies on water, energy and personnel. Results are often presented
in a rather obscure form to the public that prevents desalinization development from properly
solving real water issues, or it results too in favour. Notwithstanding, poor knowledge, and to
some extent irrational political pressure, have been responsible for these shortcomings (some
plants have been installed as a political gift and they have never been exploited since there
were no provision for operation expenses) and suboptimal use of integrated water resources,
especially with poor attention to groundwater.
Current trends to desalinization of tertiary treated sewage water are gaining ground, but
some of the experimented defects have not been adequately corrected. Sometimes expectancy
is not fully supported by well-proven technologies. However, research is still active and many
difficulties should soon be resolved.

6 OPERATION CONDITIONS OF DESALINIZATION PLANTS

Since desalinization is a capital-intensive process, as are surface water developments and


long-distance water transfers, operation at nominal capacity should be on a continuous basis,
thus keeping the number of stops for maintenance and repairs to a minimum. One important
difference regarding energy production facilities, which are also capital-intensive processes, is
the storage of energy to compensate supply variability. There is a high cost attached to energy
storage. Thus, lower capital and high energy cost facilities can be used for regulation. Water
storage, on the other hand, is generally much easier and cheaper, which means that continuous
operation of desalinization plants is feasible if enough water storage capacity is added. Surface
water dams in many cases are storage facilities whose electricity production units are relatively
The potential for desalination technologies in meeting the water crisis: comments 329

cheap but long-transfer pipelines and canals however are so capital-intensive, that projects must
generally look for full continuous use, even if in many cases they are also energy intensive.
The most difficult water storage conditions appear to be in densely populated coastal areas
due to urban pressure and the scarcity of adequate sites. The capacity for a one month supply
volume may contribute less than 0.03 a/m3 to water cost, except in very extreme situations. This
means that the continuous operation of desalinization facilities is both advisable and feasible,
and that the storage facilities required will normally have no important effect on the economic
competition among desalinization facilities or other alternatives for freshwater supply.
The conclusion is less clear for a dual purpose facility (water and electricity), depending
on the conditions for electricity production. This may be just one more cause for the difficulty
in managing these types of facilities.

7 ENVIRONMENTAL ASPECTS

There are some environmental aspects to be taken into consideration related to desalinization.
Some water managers and environmentalists are concerned about the negative results of
abstracting brackish or saline groundwater to feed desalinization plants, arguing the risk
this may have on fresh groundwater resources in the exploited aquifer. In most cases these
concerns are not real (Custodio & Bruggeman, 1987). If very simple precautions were taken,
then what would be produced would be a protection of freshwater resources against further
salinization. This is one of the hydromyths (Custodio, 2005) that must be overcome. When
brackish water is abstracted, only a fraction of the freshwater resources from the aquifer system
may be affected, but this flow must be considered in freshwater balances. A combination of
freshwater abstraction and brackish water development for desalinization is under study in the
Netherlands (Kooiman et al., 2005).
A much more serious concern is the disposal of return flows, which are often brines. For
coastal desalinization plants the disposal site is generally the sea. This slightly warm and
denser-than-seawater brine not only is a threat to the raw water supply for the plant if some
of it gets back into the feed system, but it is also a hazard for marine life, especially bottom
species and sensitive posidonia prairies, which are important habitats in some Mediterranean
sea submerged platforms. Control can be obtained through longer outfall pipes (up to 5 km and
more) and increased diffusion, thus swelling the cost of desalinization. It is assumed here that
hazardous chemicals are being disposed of safely through a separate system, or the effluent
is being treated before disposal. For example, in electrodyalysis brines high free chlorine
concentrations come from the process.
For inland plants using brackish or saline groundwater brine disposal may become a major
problem, risking aquifer saline pollution and land degradation. In fact when environmental
authorities fail to carry out controls, this is a real problem. It is even more acute for small
reverse osmosis plants, where the brine is disposed of in some cases out in the back-yard or
in old wells, or injected into the ground without a sound practice and a careful study of future
hazards.
Brine disposal requires brine-ducts to collect and convey the return waters to the disposal
site. This is expensive and prone to leakage and failures as well as requiring good surveillance
and maintenance. Disposal may be carried out in the sea, with the aforementioned restric-
tions, if the coast is not too far away. Otherwise, brines must be evaporated in natural or
330 E. Custodio

forced-evaporation ponds, and the precipitated salts must be contained to prevent future dis-
solution. If there are certain favourable circumstances and it can be carried out safely, after
some treatment brines may also be injected into deep saline aquifers. This, however, is not a
common situation. This all adds up to desalinization costs.
Disposal is easier when there is a relatively small return flow of sewage or used water from
desalinization plants, but the problem is still worthy of consideration, especially if there is an
accumulation of some harmful substances, meaning that resulting sludges and cartridges must
be disposed in a security landfill. The waste problem here is not only for the desalinization
plant itself but also for the pre-treatment stages.
The environmental concerns of desalinization adding to global problems through the gen-
eration of green-house gases and particles are real for the direct or indirect consumption of
fossil fuels. But other water supply alternatives may also be energy intensive such as the power
required for long water transportation or the prior treatment of used water.
Furthermore, energy spent on the construction of large water facilities (canals, pipes,
drainages, earth displacement, etc.) and the production of associated construction materials
(cement, iron, etc.) should also be taken into consideration in energy balances.

8 SOCIAL IMPACT OF DESALINIZATION

No doubt desalinization is rapidly expanding and will continue to expand solving real water
supply problems. Nevertheless, the cost of the water produced is limited to the demands of
those that can afford its price. This means relatively rich towns, tourist areas, non water-
intensive industries and special cash crops. Poor citizens and farmers will be left out of the
benefits unless important subsidies are applied or unless the water made available is only
for subsistence, or for strict potable uses, say a few litres per day per person. Any remaining
water needs may have to be supplied with less quality water (brackish, with some inconvenient
and non-toxic constituent, reused). For this purpose small reverse osmosis plants have been
installed in small urban communities, for instance in some inland villages in Southeast Spain,
in some convenient public emplacement for local inhabitants having easy access to it. Brine
disposal may often become a neglected problem, and a future nightmare.

9 CONCLUSIONS

The excellent chapter by Uche et al. (this volume) is complemented here with comments to
point out that desalinization is currently a significantly well-developed technology that has
achieved important cost reductions. Even though improvements can still be expected, they are
not expected to be dramatic. Therefore, the water produced by it will remain costly and energy
intensive, even though other alternative water supply resources may be as equally costly and
energy intensive.
Desalinization plants must be operated continuously, which means that enough storage
volume of water has to be provided. Artificial recharge of aquifers is an interesting alternative
for this type of storage in coastal areas.
Desalinization produces brines that must be safely disposed of, adding significantly to the
water the cost.
The potential for desalination technologies in meeting the water crisis: comments 331

Brackish groundwater is an important and often cheaper resource for desalinization;


however, the duration of resources and the brine disposal must be considered carefully.
Desalinization is a proven technology and its processes are well known. This means that
produced water costs can be calculated with some degree of confidence once the economical
parameters have been defined. These economical parameters, however, may vary from case
to case and may have been modified through social and political circumstances and pressure.
This is often done in a non-explicit manner resulting in cost scenarios that cannot be easily
compared. The result is misinformation and difficulties in making a good selection among
competing alternatives. Citizens and even policy-makers lack often reliable information. This
situation fuel obscure political manoeuvrings and the rise of a water fundamentalism that
unnecessarily complicate sound water management and the fulfilling reasonable social goals.

ACKNOWLEDGEMENTS

The author thanks the Marcelino Botín Foundation and particularly Dr. M. Ramón Llamas
for his invitation to prepare these comments. Mr. Argimiro Huerga and the General Director’s
secretariat, of the Geological Survey of Spain, have take care of the finishing of this chapter.
The ideas expressed in these comments are the author’s and not necessarily that of the organi-
zation he is linked with. With this paper I would like to pay a tribute to my friend and universal
developer of Ecology as a science, Dr. Ramón Margalef, who passed away in Spring 2004.
Humans, and especially those dealing with water owe him and this confidence in God’s role,
many guidelines and basic knowledge.

REFERENCES

Custodio, E. (2005). Myths about seawater intrusion in coastal aquifers. In: Groundwater and Saline
Intrusion (18 SWIM, Cartagena 2004). Hidrogeologia y Aguas Subterráneas 15, Instituto Geológico
y Minero de España, Madrid: 599–608.
Custodio, E. & Bruggeman, G.A. (1987). Groundwater problems in coastal areas. Studies and Reports
in Hydrology, 45: 1–576. UNESCO, Paris, France.
Custodio, E. & Llamas, M.R. (1976). Hidrología subterránea (Groundwater hydrology). 2 vol.: 1–2350.
Ediciones Omega, Barcelona, Spain.
Kooiman, J.W.; Stuyfzand, P.J.; Maas, C. & Kappelhof, J.W.N.M. (2005). Pumping brackish groundwater
to prepare drinking water and keep salinizing wells fresh: a feasibility study. In: Groundwater and
Saline Intrusion (18 SWIM, Cartagena 2004). Hidrogeologia y Aguas Subterráneas 15, Instituto
Geológico y Minero de España, Madrid: 625–635.
Llamas, M.R. & Custodio, E. (2003). Intensive use of groundwater: a new situation which demands
proactive action. In: M.R. Llamas & E. Custodio (eds.), Intensive Use of Groundwater: Challenges
and Opportunities. Balkema: 13–31.
Merrett, S. (2004). The demand for water: four interpretations. Water International, 29(1): 27–29.
Semiat, R. (2000). Desalination: present and future. Water International, 25(1): 54–65.
Shiklomanov, V. (1998). World water resources: a new appraisal assessment for the 21st century.
UNESCO, Paris, France.
Author index

Aguirre, M.S. 93 Del Porto, D. 285 Mukherji, A. 181


Allan, J.A. 131
Arrojo, P. 119 Frederick, K.D. 105 Ramirez-Vallejo, J. 151
Asano, T. 261 Rogers, P. 3
García Bouzas, F. 235
García Novo, F. 235 Schlager, E. 43
Bergkamp, G. 253
Hanemann, W.M. 61 Serra, L. 297
Bhatia, M. 197
Sullivan, C.A. 221
Bhatia, R. 197
Llamas, M.R. 163
López-Gunn, E. 43 Uche, J. 297
Cosgrove, W.J. 37
Custodio, E. 323 Martínez-Santos, P. 163 Valero, A. 297

You might also like